GREEN’S FUNCTIONS FOR VARIANTS OF THE SCHRAMM-LOEWNER EVOLUTION By Benjamin Mackey A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics—Doctor of Philosophy 2017 ABSTRACT GREEN’S FUNCTIONS FOR VARIANTS OF THE SCHRAMM-LOEWNER EVOLUTION By Benjamin Mackey We prove upper bounds for the probability that a radial SLEκ curve comes within specified radii of n different points in the unit disc. Using this estimate, we then prove a similar upper bound for the probability that a whole-plane SLEκ passes near any n points in the complex plane. We then use these estimates to show that the lower Minkowski content of both the radial and whole-plane SLEκ traces has finite moments of any order. For κ ≤ 4, the reverse flow of the Loewner equation driven by √ κBt generates a random continuous function φ : R+ → R+ called the conformal welding. In studying backward SLE, this plays the roll of the global random object, rather than the SLE trace. Given any x, y > 0 we use the Girsanov theorem to construct a family of probability measures, depending on some parameters, under which the conformal welding satisfies φ(x) = y almost surely. For one such law, we prove a one-point estimate for the backward SLE welding and show how it coincides with the Green’s function. In another case, we decompose the law of the welding conditioned to pass through (x, y) into two pieces. Using this decomposition, we integrate this law over a set U ⊂ [0, ∞) × [0, ∞) to get a new measure on weldings which is absolutely continuous with respect to the original backward SLE welding. Moreover, the Radon-Nikodym derivative is given by the capacity time that the graph of φ spends in U . In the last chapter, we study a generalization of the chordal Loewner equation called chordal measure driven Loewner evolution. We show existence of a solution to the equation, and a one-to-one correspondence between the appropriate measures and all continuously growing families of H-hulls. In [19], the notion of measure driven Loewner evolution was first introduced in the radial setting, and a similar theorem was proven. This result is pure complex analysis, without any reference to probability theory. I would like to dedicate this thesis to my wonderful wife, Lauren Conway, who has never waivered in her support of and confidence in me. I could not have written this without her standing by my side.“Best of wives and best of women.” iv ACKNOWLEDGMENTS I would like to expresss my sincerest gratitude for my advisor, professor Dapeng Zhan, for introducing me to the subject of SLE and for continuously providing support for my research. His guidance in selecting research problems on which to work has been invaluable, as has his perpetual willingness to comment on my progress and to steer my efforts into a productive direction. I also would like to thank him for his essential assistance in editting my thesis into its current form. I would also like to thank the members of my committee, professors Shlomo Levantal, Ignacio Uriate-Tuero, and Yimin Xiao, for the taking the time to read and comment on my research. I am also grateful that they have taken the time to serve on my committee during the summer semester. I particularly want to thank professor Ignacio Uriarte-Tuero for serving as a constant figure in my education here at Michigan State University and always expressing an interest in my work. His real analysis course pushed me harder than any other course I have ever taken or will take, and I cannot quantify how much I gained from the experience. Lastly, I would like to thank my friends and family for their constant support during my time in graduate school. Tyler Bongers, Guillermo Rey, Reshma Menon, Emily Olson, Rodrigo Matos, Leonardo Abbrescia, Rami Fakhry, Chelsea Law, David Zach, Matt Alexander, Mark Bissler, Isaac Defrain, and many more have worked with me, run with me, rock climbed with me, or generally kept me sane during my time at Michigan State University. My wife, parents, brother, and in-laws have also all been incredibly accomidating of my frequent absences and interstate travel as I work on this dissertation. v TABLE OF CONTENTS Chapter 1 Introduction . . . . . . . . . . . . . . 1.1 Description of SLE . . . . . . . . . . . . . . . 1.2 Definitions of variants of SLE . . . . . . . . . 1.2.1 Hulls in the complex plane . . . . . . . 1.2.2 Chordal Loewner evolution in H . . . . 1.2.3 Radial Loewner evolution in D . . . . . 1.2.4 Whole-plane Loewner evolution in C . 1.2.5 Backward Loewner evolution . . . . . . 1.3 Natural parametrization and Green’s functions 1.4 Generalizations of the Loewner equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 2 Multipoint estimates for radial and whole-plane 2.1 Statement of results . . . . . . . . . . . . . . . . . . . . . . . 2.2 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Crosscuts and prime ends . . . . . . . . . . . . . . . 2.2.2 The covering space H∗ for D . . . . . . . . . . . . . . 2.2.3 Extremal length and conformal transformations . . . 2.3 Interior and boundary estimates . . . . . . . . . . . . . . . . 2.4 Components of crosscuts . . . . . . . . . . . . . . . . . . . . 2.5 Concentric circles . . . . . . . . . . . . . . . . . . . . . . . . 2.6 Main theorems . . . . . . . . . . . . . . . . . . . . . . . . . 2.7 Technical lemmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 4 4 6 7 7 8 9 14 SLE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 17 19 19 21 24 27 35 43 50 58 Chapter 3 Decomposition of backward SLE in the capacity parametrization 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Processes with a random lifetime . . . . . . . . . . . . . . . . . . . . . . . . 3.3 Glossary of notation for Ito’s formula calculations . . . . . . . . . . . . . . . 3.4 Generalized Green’s functions . . . . . . . . . . . . . . . . . . . . . . . . . . 3.5 Capacity parametrization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 One-point estimate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Chapter 4 Measure driven chordal Loewner evolution 4.1 Theorem statement and definitions . . . . . . . . . . . 4.2 Preliminary results . . . . . . . . . . . . . . . . . . . . 4.3 A disintegration lemma . . . . . . . . . . . . . . . . . . 4.4 Elementary calculations . . . . . . . . . . . . . . . . . 4.5 Proof of existence . . . . . . . . . . . . . . . . . . . . . 4.6 Growing Hulls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 63 65 71 72 81 93 97 97 99 102 108 115 121 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 vi Chapter 1 Introduction 1.1 Description of SLE This thesis is concerned with the study of conformally invariant randomly growing processes in subdomains of the complex plane. The main object is the Schramm-Loewner evolution, or SLE, which was introduced by Oded Schramm in [29] as a candidate for several lattice models in statistical physics. The processes all arise from solving a variant of the Loewner differential equation driven by time scaled Brownian motion. In all cases, there is a domain D ⊂ C and a random curve γ contained in the closure of D starting and ending at fixed points. To motivate the definition, we will informally describe the loop erased random walk in a domain from a to b. Let D ⊂ C be a simply connected domain, and let δ > 0 be a small number. Consider the lattice Dδ = D ∩ δZ2 , and let γδ be the path of a simple random walk on Dδ starting at the point (nearest to) a ∈ ∂D, and stopped when it hits a point (nearest to) b in D, the closure of D. The path γδ can be turned into a simple path γδLE by removing all of the loops. Is there is a random path γ so that γδLE → γ in law as δ → 0? SLE provides an answer, which is that the scaling limit is an SLE curve. SLE has a parameter κ > 0 on which the behavior of the SLEκ path depends. For the loop erased random walk, SLE2 has been proven to be the scaling limit for the loop erased 1 random walk [13]. Several other lattice models have proven to converge to SLEκ for other values of κ. SLE3 and SLE16/3 are the scaling limits for Ising model interfaces [35], SLE6 is the scaling limit of the critical percolation explorer curve [34], SLE4 is the scaling limit of the harmonic explorer [30], and SLE8 is the scaling limit of the uniform spanning tree Peano curve [13]. There are several varieties of SLE which differ based on where they begin and end, and each variety has a standard domain in which it is easier to define and study. The two primary types are called chordal SLEκ , which begins and ends at fixed boundary ponts, and radial SLEκ , which begins at a fixed boundary point and ends at a fixed interior point. Making this more precise, fix a simply connected domain D, and fix a point a ∈ ∂D (technically, a should be a prime end of D, but we will not define that now), and let b ∈ D. If b ∈ ∂D (or b is a prime end of D), then chordal SLEκ in D from a to b is a random path in D which begins at a and ends at b almost surely. The standard domain for studying chordal SLE is the upper half-plane H = {z ∈ C : Im(z) > 0}, starting at 0 and ending at ∞, where ∞ can seen as a point in the Riemann sphere. If b is in the interior of D, then radial SLEκ in D from a to b is a random path in D which begins at a and ends at b almost surely. The standard domain for radial SLEκ is the unit disc D = {z ∈ C : |z| < 1}, starting at 1 and ending at 0. Once the SLE path is defined in its standard domain, in either case, it can be defined in any simply connected domain D. To define chordal SLEκ in a simply connected domain D from a to b, let φ : H → D be a conformal tranformation with φ(0) = a and φ(∞) = b. Then if µκ is the probability measure on paths in H which is the law of the standard chordal SLEκ , then the law of SLEκ in D from a to b is given by the pushforward meaure induced by φ on µκ . 2 This describes conformal invarience. Schramm proved that the only processes which have conformal invarience and the domain Markov property are SLE processes. These properties are described more precisely as: • (Conformal invarience) If γ is an SLEκ path in D from a to b, and φ : D → D is a conformal transformation with φ(a) = a and φ(b) = b , then φ(γ) is a time changed SLEκ path in D from a to b . • (Domain Markov property) Let T be a stopping time with respect to the filtration generated by an SLEκ path in a domain D from a to b. For any t > 0, Let Dt be the component of D\γ[0, t] which contains the target point b. Then if γ T (t) = γ(T + t), after conditioning on γ[0, T ], the path γ T is an SLEκ path in DT from γ(T ) to b. The laws of chordal SLE and radial SLE are not absolutely continuous with each other, but there is local absolute continuity [31]. If γ is radial SLEκ in D from 1 to 0, let > 0 and T = inf{t > 0 : |γ(t)| = }, which is finite a.s.. Then radial SLEκ restricted to time t ≤ T is absolutely continuous with respect to a stopped chordal SLEκ up to a time change, so the paths share many of the same local properties. For example, in each case, the Hausdorff dimension of the SLEκ path is d = min{1 + κ/8, 2} [3]. Also, the path has two phase transitions as κ varies [27]. We state the transitions when γ is the chordal SLEκ path in H from 0 to ∞. For 0 < κ ≤ 4, γ(0, ∞) is a simple path which stays in H. For κ ≥ 8, γ is a space filling path. That is, γ(0, ∞) = H almost surely. For 4 < κ < 8, γ has both boundary intersections and self intersections. If κ < 8, for any z ∈ H, we have P[z ∈ γ] = 0. This is proven in [27] reducing the SLEκ maps started on the boundary to the study of Bessel processes. Note that if κ ≤ 4, the domain DT , as described in the domain Markov property, is 3 exactly equal to D\γ[0, T ], since the path is simple and has no boundary intersections. For κ > 4, DT the piece of the domain which has not yet been cut off from the endpoint by the path. Another variety of SLE is called whole-plane SLE, which is a path in C which grows from 0 to ∞. It can be seen as a limit of radial SLEκ paths in large discs from the boundary to ˆ to any other b ∈ C, ˆ where C ˆ is 0. This can be extended to whole-plane SLE from any a ∈ C the extended complex plane, by applying a M¨obius transformation which takes 0 to a and ˆ ∞ to b, which gives an SLE path connecting any two interior points of C. The domain Markov property has a slightly different form for whole-plane SLEκ . If γ ∗ is the whole-plane SLEκ path, then γ ∗ is a path γ ∗ : (−∞, ∞) → C. For any stopping time T , the path γ T (t) = γ ∗ (T + t), conditioned on γ ∗ (−∞, T ], is a a radial SLEκ in C\DT , where DT is the unbounded component of C\γ ∗ (−∞, T ]. 1.2 Definitions of variants of SLE There are several varieties of the Loewner equation. First, we describe the types of hulls which correspond to each Loewner equation. We then briefly discuss the chordal equation, then introduce radial and whole-plane Loewner equations. Complete details can be found in [7]. We also introduce the reverse Loewner equation and define backward SLE. 1.2.1 Hulls in the complex plane Let H = {z ∈ C : Im(z) > 0} be the upper half-plane. A set K ⊂ H is called an H-hull if K is relatively compact in H and H\K is simply connected. Then there is a unique conformal map gK : H\K → H with gK (z) = z + c + O(z −2 ) as z → ∞. The constant c is called z 4 the half-plane capacity, and is denoted by hcap(K), and it is equal to 0 if and only if K is empty. We call a domain H ⊂ H and H-domain if H = H\K for some H-hull K. Half-plane capacity is monotonic, in that if K1 ⊂ K2 are both H-hulls, we have hcap(K1 ) ≤ hcap(K2 ). Moreover, if K1 ⊂ K2 are both H-hulls, we can define a new H-hull K2 /K1 := gK1 (K2 \K1 ), called the quotient hull. In [26], it has been shown that hcap(K2 ) = hcap(K1 )+ hcap(K2 /K1 ). Therefore, if K1 is properly contained in K2 , it follows that hcap(K1 ) < hcap(K2 ). Half-plane capacity also satisfies the scaling rule hcap(rK + a) = r2 hcap(K) for any a, r ∈ R. In [26], the authors define the support of an H-hull K and prove the following results. ˆ = K ∪ K ∪ BK . Let BK = K ∩ R, and let K = {z : z ∈ K}. Define the double of K by K ˆ = C\SK , for a compact SK ⊂ R, which we will call the support of K. Then Then gK (C\K) −1 defined in H can be extended conformally to C\S by the Schwarz the inverse fK := gK K reflection principle, but no further. If K only has one component, then SK is an interval, ∗ denote the convex hull of S , but in general K can consist of several components. Let SK K ∗ is a compact interval. If K ⊂ K , then S ∗ ⊂ S ∗ . so that SK 1 2 K1 K2 Let D = {z ∈ C : |z| < 1} be the unit disc. A D-hull is a set K ⊂ D which is relatively closed in D, 0 ∈ / K, and D\K is simply connected. By the Riemann mapping theorem, there is a unique conformal map gK : D\K → D with gK (0) = 0 and gK (0) > 0. The D-capacity of K is defined by dcap(K) = ln(gK (0)). A set D ⊂ D is called a D-domain if D = D\K for a D-hull K. D-capacity can be readily shown to be monotonic with respect to inclusion. If K1 ⊂ K2 and the quotient hull K2 /K1 := gK1 (K2 \K1 ), then gK2 = gK /K ◦ gK1 , and the chain rule 2 1 implies that gK (0) = gK /K (0)gK (0). Taking the natural logarithm of both sides yields 2 1 2 1 the monotonicity. 5 A compact hull in C is a set K ⊂ C such that 0 ∈ K, K is connected and compact, and C\K is connected. Then there is a unique conformal map FK : C\D → C\K with limz→∞ FK (z)/z > 0. Looking at C\K under the inversion z → 1/z gives a simply connected domain U containing 0, for which there is a unique conformal Riemann map fK : D → U with fK (0) > 0. Then FK (z) = 1/fK (1/z), and the logarithmic capacity of K is defined by cap(K) = log(fK (0)). 1.2.2 Chordal Loewner evolution in H The chordal Loewner equation describes the growth from a boundary point to another boundary point. The standardized domain is the upper half-plane H = {z ∈ C : Im(z) > 0}, and the chordal Loewner equation driven by a real valued λ ∈ C[0, ∞) at z ∈ H is ∂t gt (z) = 2 , gt (z) − λt g0 (z) = z. (1.1) If τz = inf{t > 0 : Im(gt (z)) = 0} is the lifetime of (1.1) at z, define Kt = {z ∈ H : τz ≤ t}. Then Kt is an H-hull with half-plane capacity hcap(Kt ) = 2t, and gt = gKt is the conformal map sending H\Kt to H. If κ > 0 and Bt is a standard one dimensional Brownian motion, then (1.1) driven λt = √ κBt is called chordal SLEκ . In [27], it is proven that there is a path γ : [0, ∞) → H called the chordal SLEκ trace, also called the path or curve, which grows from 0 to ∞. Then for each t > 0, H\Kt is the unbounded component of H\γ[0, t]. 6 1.2.3 Radial Loewner evolution in D Given any real valued continuous function λ ∈ C[0, ∞), the radial Loewner equation driven by λ at z ∈ D, ∂t gt (z) = gt (z) eiλ(t) + gt (z) eiλ(t) − gt (z) , g0 (z) = z. (1.2) If τz = inf{t > 0 : |gt (z)| = 1} is the lifetime of (1.2) at z ∈ D and Kt = {z ∈ D : τz ≤ t}, then Kt is a D-hull with dcap(Kt ) = t. For κ > 0, the radial SLEκ process is the solution (1.2) for λ(t) = √ κBt , where Bt is standard one dimensional Brownian motion. Similarly to the chordal case, there is a a radial trace γ : [0, ∞) → D so that D\Kt is the component of D\γ[0, t] containing 0 with γ(0) = 1 and γ(∞) = 0. The radial SLEκ trace has the same phrase transitions as the chordal SLEκ and the same dimension. The radial SLEκ process can be studied by looking at the covering space for D under the exponential map eiz , which is the cylinder H∗ = {[z]∼ : z ∈ H}, where z ∼ w if z − w ∈ 2πZ. This space will be discussed in more detail in Section 2.2.2. 1.2.4 Whole-plane Loewner evolution in C If λ ∈ C(−∞, ∞) is real valued and z ∈ C\{0}, the whole-plane Loewner equation driven by λ is e−iλt + gt∗ (z) ∂t gt∗ (z) = gt∗ (z) −iλ , and lim et gt∗ (z) = z. t→−∞ e t − gt∗ (z) (1.3) This can be interpreted as a radial Loewner equation driven by −λ started from t = −∞. For each t ∈ (−∞, ∞), Kt = {z ∈ C : τz ≤ t} is a compact hull in C with logarithmic capacity cap(Kt ) = t, and gt∗ : C\Kt → C\D conformally. To define whole-plane SLEκ , we need to define two sided Brownian motion. Let Bt1 and 7 Bt2 be independent standard one dimensional Brownian motions, and let Y be an independent √ √ uniform [−π/ κ, π/ κ] random variable. Construct a two sided Brownian motion B : R → 2 + Y if t < 0. Then g ∗ is a whole-plane SLE if it R by Bt = Bt1 + Y if t ≥ 0, and Bt = B−t κ t √ solves (1.3) with λt = κBt . There is a curve γ ∗ : (−∞, ∞) → C called the whole-plane SLEκ trace. For each t ∈ R, C\Kt is the unbounded component of C\γ ∗ (−∞, t]. Also, limt→−∞ γ ∗ (t) = 0 and limt→∞ γ ∗ (t) = ∞, and so the whole-plane SLEκ curve grows from 0 to ∞ in the entire complex plane. Since at any time s ∈ R the process (Bt )t≥s is a Brownian motion with a random start time, it follows that γ ∗ has the same phrase transitions and dimension depending on κ as the chordal and radial processes. 1.2.5 Backward Loewner evolution The backward Loewner equation started at z ∈ H driven by λt ∈ C[0, T ) is ∂t ft (z) = −2 , ft (z) − λt f0 (z) = z. (1.4) The process (ft )t≤T is called the backward Loewner process driven by λ. Suppose solving equation (1.1) with the function t → λ(t0 − t) generates a forward Loewner trace t → γt0 (t0 − t) for each t0 ≤ T . Then we say that λ generates a family of backward Loewner traces (γt0 )t0 0, and let λt = √ κBt . Then the process given by solving (1.4) is called 8 the backward SLEκ process, and will be denoted by BSLEκ . For a fixed T < ∞, the trace γT − λT has the same law as a forward SLEκ trace on [0, T ]. So, to study the SLE trace at a finite time, it suffices to study backward SLE, where the Loewner equation may be easier to work with. However, the backward SLE traces are not a good global object to study, since the images γT [0, T ] evolve over time, rather than grow from the tip. If the traces are simple, there is another object called the conformal welding which serves as global object to study. For κ ≤ 4, the SLEκ traces are simple, and so the BSLEκ process generates a random function φ : R+ → R+ called the conformal welding. The conformal maps ft can be extended to the boundary of H. For any x ∈ R, let τx denote the lifetime of the backward Loewner equation, which is finite almost surely. For x, y > 0, the welding is defined by φ(x) = y if τx = τ−y , which is true if and only if ft (x) = ft (−y) for some t > 0. Then φ is a random monotonic function on R+ . This welding is the main object of study in [26], where the welding is shown to satisfy a reversability property analogous to a fundamental property of the forward SLE trace [38] [20]. We will also need to keep track of which points are welded together at a given capacity time t. To this end, let bt = sup{x > 0 : τx ≤ t} and at = sup{y > 0 : τ−y ≤ t}, so that ft (bt ) = ft (−at ) = λt for every t. Define Φ(t) = (bt , at ), which we will call the welding curve. Then if Q1 = [0, ∞) × [0, ∞) is the first quadrant, Φ : [0, ∞) → Q1 is a random path whose image is the graph of the welding function φ. 1.3 Natural parametrization and Green’s functions The lattice models which converge to SLE paths converge not only as sets, but also as continuous functions. To make that conclusion, the lattice model paths and the SLE paths 9 must be parametrized in a particular way. By construction, SLE is parametrized so that the capacity of the path grows as a constant rate. The lattice model paths which converge to SLE must also use the capacity parametrization, which is not the natural way to run the model. It is desirable to run the models so that each discrete segment takes the same length of time. The most intuitive approach is to simulaneously run the SLE path parametrized by length, but since the fractal dimension of SLEκ is d = min{1 + κ/8, 2} > 1, the length of the SLEκ trace is always infinite. Recently, there has been a program to develop a d-dimensional measurement of length for chordal SLEκ which can be used to study convergence. In [12], the Doob-Meyer theorem was used to create an increasing process related to the path which was called the natural parametrization, or natural length, for κ < 5.021.... It was conjectured to coincide with some d-dimensional measurment. In [17], the Doob-Meyer construction was extended to all κ < 8. In [22], it was proven that the Hausdorff measure of the path is 0 almost surely. Instead, the Minkowski content of the path has been proven to be the correct candidate for the natural parametrization. The d-dimensional Minkowski content of a set E ⊂ C is defined by Contd (E) = lim rd−2 Area{z ∈ C : dist(z, E) < r}, r→0 provided that the limit exists. In [9], it was proven that the d-dimensional Minkowski content of the path exists almost surely, and that it differs from the Doob-Meyer construction by a multiplicative constant. The first proof of convergence in the natural parametrization is [14], where it is proven that the loop erased random walk converges to SLE2 when parametrized by the Minkowski content. One of the main tools for studying the natural parametrization is called the Green’s 10 function, which gives the normalized probability that the SLE path passes through a point. For κ ≥ 8, P[z ∈ γ] = 1 for all z, so the question is not interesting. For κ ∈ (0, 8), P[z ∈ γ] = 0 implies that the probability P[dist(z, γ) < r] should decay as r → 0. Moreover, if the dimension of the path is d, the probability should decay as r2−d . The Green’s function is defined to be G(z) = lim rd−2 P [dist(z, γ) < r] , r→0 provided this limit exists. In the chordal case, this was first studied with the event {radH\γ (z) < r} rather than distance, where radD (z) is defined to be the conformal radius of D at z (defined in Section 2.2.2). In [27], it was observed that for chordal SLEκ and any z ∈ H, the process Mt (z) = |gt (z)|2−d G(gt (z) − √ κBt ) should be a local martingale. To construct the natural parametrization in [12], the authors weighted the probability measure by Mt (z) and used the Girsanov theorem to construct what is called two-sided radial SLEκ through z, which is chordal SLEκ conditioned to pass through z. In order to further study the Minkowski content, the two point Green’s function was proven to exist and analyzed [9] [10] [16]. In [23] and [24], the Green’s function for any finite number of points was proven to exist for chordal SLEκ . That is, n rkd−2 P ∩n k=1 dist(zk , γ) ≤ r G(z1 , . . . , zn ) = lim r→0 k=1 was proven to exist, where r = (r1 , . . . , rn ) and z1 , . . . , zn ∈ H are arbitrary. The upper bound for the multipoint Green’s function found in [23] is used to prove that the Minkowski 11 content has finite moments of all orders. In [37], the construction of two-sided radial SLE using the Girsanov theorem is generalized. For ρ ∈ R and z ∈ H, functions Gρ (z) are found so that if ρ Mt (z) = |gt (z)|p Gρ (gt (z) − √ κBt ) ρ for an appropriate power p = p(ρ), then Mt (z) is a local martingale which can be used to ρ ρ create a measure Pz under which Pz [z ∈ γ] = 1. A few special cases of these generalized Green’s functions are used to prove decompositions of the SLEκ path. In each case, the ρ ρ measures Pz are averaged over z ∈ U weighted by Gρ (z) to generate a new measure PU , which is absolutely continuous with respect to the original measure P. The Radon-Nikodym derivatives for three of these constructions are found, and are shown to be the Minkowski content of γ ∩ U , the capacity time spent by γ in U , and the analogue of the natural parametrization for the boundary intersection of γ in U ⊂ R for κ ∈ (4, 8). The result for Minkowski content extends the result in [5] which was proven for κ ≤ 4. These decomposition theorems have been used to define SLE loop measures [41]. The Green’s function for radial SLEκ is not as well understood as the Green’s function for chordal SLEκ . In [2], the conformal radius version of the one-point Green’s function is proven to exist for κ ∈ (0, 8), though an exact form is only found for κ = 4. In Chapter 2, we establish an upper bound for the probability that radial SLEκ in D from 1 to 0 passes near any finite collection of points in D. This should serve as the first step towards proving the multipoint Green’s function for radial SLEκ exists. The strategy is to move from the covering space H∗ of D under the exponential map eiz , where H∗ is the upper half-plane H under the equivalence relation z ∼ w if z −w ∈ 2πZ. Here, one-point estimates are developed 12 in [2] for both the boundary and interior case, which we adapt into appropriately general and conformally invarient estimates. After establishing a sufficiently robust one-point estimate for radial SLEκ , we follow the strategy in [23] to extend the estimate to multiple points. The Minkowski content of radial SLEκ and whole-plane SLEκ have not yet been proven to exist, but the lower content Contd (E) = lim inf rd−2 Area{z ∈ C : dist(z, γ) ≤ r} r→0 always exists. Using the multipoint estimate for radial SLEκ we develop, we prove that Contd (γ) has finite moments of all orders. If the Minkowski content were proven to exist, the same proof would show that Contd (γ) has finite moments of all orders. Using the reversibility of whole-plane SLEκ [40] [21] and the domain Markov property, we then establish a similar upper bound for the probability that a whole-plane SLEκ passes near any finite collection of points using the estimate for radial SLEκ . Using this estimate and a proof similar to the radial case, we then prove that the lower Minkowski content of a whole-plane SLEκ path in any compact set has finite moments of any order. In Chapter 3, we establish a theory of Green’s functions for the backward SLEκ welding for κ ≤ 4. Given any two points x, y > 0 we establish estimates for the probability of the event that x and y are almost welded together in a way analogous to the estimate of the probability that an SLEκ path passes near a point. Normalizing this estimate yields a function G(x, y) which we call a backward SLEκ Green’s function. Moreover, we observed that the process Mt (x, y) = G(ft (x) − √ √ κBt , κBt − ft (−y)) is a local martingale, and weighting the original probability measure by Mt (x, y) gives a 13 measure Px,y under which φ(x) = y almost surely. Following the strategy in [37], we generalize this construction to get a family of functions Ga,b (x, y) which serve as generalized Green’s functions for the BSLEκ welding. For any a, b, we find powers p, q so that a,b Mt (x, y) = ft (x)p ft (−y)q Ga,b (ft (x) − is a local martingale. If a, b ≤ − √ √ κBt , κBt − ft (−y)) 4+κ , using this local martingale with the Girsanov theorem 2 a,b yields a measure Px,y under which φ(x) = y almost surely. For a = b = −4, we show that this measure gives insight into the BSLEκ welding in the capacity parametrization in a way similar to what has been proven for forward chordal SLEκ . Given any U ⊂ [0, ∞)×[0, ∞), we average the law of the welding conditioned to pass through −4,−4 U by integrating Px,y against G−4,−4 (x, y) over U . This yields a decomposition theorem similar to those proven in [37] for the welding, and gives a measure PU which is absolutely continuous with respect to the original probability measure. The resulting Radon-Nikodym derivative is then shown to be the capacity time spent by the graph of the welding φ in U . 1.4 Generalizations of the Loewner equation When Charles Loewner first introduced his differential equation in 1923 [18], it was a tool used to study the Bieberbach conjecture. His version of the equation was the radial Loewner equation, and the setting was entirely deterministic. The strategy was to solve (1.2) using real valued continuous functions to study slit-domains as a way to answer extremal questions about conformal maps in the unit disc. In [29], Oded Schramm recontextualized Loewner’s 14 ideas in order to study scaling limits of discrete random processes by using a Brownian motion as the driving function. Along with introducing randomness, Schramm also introduced the chordal Loewner ODE to study growth from a boundary point to another boundary point. In the text [7], Lawler introduces generalized versions of both the chordal and radial Loewner equation. Rather than being driven by a continuous function λt , the equations he introduces are driven by a family of locally bounded Borel measures {µt }t≥0 on R (or ∂D in the radial case) so that t → µt is weakly continuous. In the chordal case, the Loewner equation driven by {µt }t≥0 is ∂t gt (z) = 1 dµt (u), R gt (z) − u g0 (z) = z. Then for each t > 0, gt : H\Kt → H is conformal for some increasing family of H-hulls (Kt )t≥0 with hcap(Kt ) = 0t µs (R)ds. This family of measures can be reparametrized so that the corresponding family of hulls Kt satisfy hcap(Kt ) = 2t. For a real valued continuous function λ, the regular chordal Loewner equation driven by λ is the above equation with µt = δλt , where δx is the point mass measure at x. In [19], Miller and Sheffield introduce a further generalization in the radial case which they called measure driven Loewner evolution. They establish existence and uniqueness in the radial setting, and then prove a one-to-one correspondence between families of D-hulls and solutions to the measure driven Loewner evolution. Moreover, convergence on hulls was shown to be equivalent to convergence of the corresponding measures. In Chapter 4, we introduce measure driven chordal Loewner evolution. For a measure µ on [0, ∞) × R with the appropriate assumptions, we show existence and uniqueness of the 15 solution to the integral equation gt (z) − z = 2 dµ(s, u) [0,t]×R gs (z) − u for all z ∈ H. For a family of measures {µt }t≥0 as in Lawler’s definition, we define a measure µ by dµ(t, u) = dµt (u)dt so that µ and {µt }t≥0 generate the same process.. If N is the family of measures which generate a measure driven chordal Loewner equation, any µ ∈ N generates a family of H-hulls (Kt )t≥0 with hcap(Kt ) = 2t so that gt = gKt . Conversely, we show that any family of H-hulls (Kt )t≥0 parametrized by capacity is the resulting family of hulls for some measure µ ∈ N . Moreover, we endow N with a topology and show that a family of measures µn → µ in N if and only if the domains H\Ktn → H\Kt in the Carath´eodory topology (to be defined in Section 4.2) for each t, which is equivalent to the conformal maps (gtn )−1 converging locally uniformly to gt−1 . 16 Chapter 2 Multipoint estimates for radial and whole-plane SLE 2.1 Statement of results The main theorem of this chapter is the multipoint estimate for radial SLEκ . Theorem 1. Fix κ ∈ (0, 8). Let γ be a radial SLEκ in D from 1 to 0, let zk ∈ D\{0} for k = 1, . . . , n, and let z0 = 1. Let yk = 1 − |zk | be the distance of each point to the boundary, and define lk = min{|zk |, |zk − 1|, |zk − z1 |, . . . , |zk − zk−1 |}. Then there exists an absolute constant Cn < ∞, depending on κ and n, so that n Pyk (rk ∧ lk ) P[∩n k=1 {dist(γ, zk ) < rk }] ≤ Cn k=1 Py (l ) k k . (2.1) The function Py (x) used in the upper bound in Theorem 1 is defined by Py (x) =     y α−(2−d) x2−d , x ≤ y    x α , , x≥y where d = 1 + κ/8 is the Hausdorff dimension of the path, and α = 8/κ − 1 is related to the boundary exponent for SLEκ . This upper bound mixes the estimates for interior points and 17 points near the boundary. Roughly speaking, if the point zk is far from the boundary, the term on the right hand side of (2.1) corresponding to zk will be (rk /lk )2−d . If zk is near the boundary of the unit disc, then the corresponding term on the right side of (2.1) is (rk /lk )α . If zk is close, but not too close, then the corresponding estimate is a mixture of the two. The following Lemma about the functions Py is Lemma 2.1 in [24], and can be proven with a case by case argument. Lemma 1. For 0 ≤ x1 < x2 , 0 ≤ y1 ≤ y2 , 0 < x, and 0 ≤ y, we have Py1 (x1 ) Py (x1 ) ≤ 2 ; Py1 (x2 ) Py2 (x2 ) x1 α Py (x1 ) ≤ ≤ x2 Py (x2 ) x1 d−2 Px2 (x1 ) ; = x2 Px2 (x2 ) y1 α−(2−d) Py1 (x) ≤ ≤ 1. y2 Py2 (x) Theorem 1 can then be used to prove a similar estimate for whole-plane SLEκ : Theorem 2. Fix κ ∈ (0, 8), and let γ ∗ be a whole-plane SLEκ trace from 0 to ∞. Let z1 , . . . , zn ∈ C\{0}. For each k = 1, . . . , n, let 0 < rk < |zk | and define lk = min{|zk |, |zk − z1 |, . . . , |zk − zk−1 |}. Then n ∗ P[∩n k=1 {dist(γ , zk ) < rk }] ≤ Cn k=1 rk ∧ lk 2−d . lk (2.2) Note that the expression of this bound is simpler than Theorem 1, since there are no boundary points with which to be concerned. Both of these estimates are used to prove the 18 following theorem about Minkowski content: Theorem 3. Fix κ ∈ (0, 8). a) Let γ be a radial SLEκ trace in D from 1 to 0. Then E[Contd (γ)n ] < ∞ for all n ∈ N. b) Let γ ∗ be a whole-plane SLEκ trace from 0 to ∞, and suppose D ⊂ C is compact. Then E[Contd (γ ∗ ∩ D)n ] < ∞ for every n ∈ N. The chapter will be organized as follows. First we review preliminary information, including information about the covering space H∗ . We also review some information about crosscuts, prime ends, and extremal length which will be used. Next, we provide one-point estimates for radial SLE in the forms which will be useful for us. We then use these one-point estimates to prove some key lemmas, followed by the proofs of the main theorems. 2.2 2.2.1 Preliminaries Crosscuts and prime ends In later sections, we will be studying the behavior of the radial SLE curve as it crosses many interior curves, creating different components of the initial domain D. We need to introduce some notation which will make it easier to distinguish which component is discussed at any point in time. This is the same framework introduced in [23]. Recall that a crosscut in a simply connected domain D ⊂ C is a simple curve ρ : (a, b) → D such that limt→a+ ρ(t) := ρ(a+ ) and limt→b− ρ(t) := ρ(b− ) both exist and are elements of the boundary of D. Then ρ lies inside D, but the endpoints do not. The endpoints ρ(a+ ) and ρ(b− ) determine prime ends for the domain. Note that if f : D → D is a conformal 19 map, then f (ρ) is a crosscut in D . More information about crosscuts and prime ends can be found in [1]. Note that if ρ is a crosscut in a domain D, then ρ divides D into two components. Even more generally, let K ⊂ D be relatively closed. Let S be either a connected subset of D\K or a prime end of D\K. We then define D(K; S) to be the component of D\K which contains S. We also introduce the symbol D∗ (K, S) = D\ (K ∪ D(K; S)), which is the union of the remaining components of D\K. This notation is useful for expressing whether K seperates points. For example, if ρ ⊂ D is a crosscut which separates two points z, w ∈ D, then D(ρ; z) = D(ρ; w). In fact, in this case, we have D(ρ; w) = D∗ (ρ; z), and D(ρ; z) = D∗ (ρ; w). Since we will be working with domains which have 0 as an interior point, and in particular will be concerned with components containing 0, we will use D(K) and D∗ (K) to represent D(K; 0) and D∗ (K; 0) respectively. Note that this is a departure from the notation in [23], where the point being suppressed was the prime end ∞. Since we are working with radial SLE rather than chordal SLE, the change is to reflect the fact that we are looking at components with respect to the endoint of the SLE curve, which in this work will be the interior point of 0. For an example of how this notation will be used, let γ be a radial SLEκ curve in D from 1 to 0, and define Dt = D(γ[0, t]). This is the component of D\γ[0, t] which contains the origin. If ρ is a crosscut in Dt , and z ∈ Dt \ρ, we will ask if z0 ∈ Dt (ρ). That is, does ρ seperate z0 from 0 in Dt ? Suppose the circle ξ = {|z − z0 | = r} is contained in Dt \ρ. If ξ doesn’t enclose 0, then D∗ (ξ) = {|z − z0 | < r}. Then D∗ (ξ) ⊂ Dt∗ (ρ) means the circle ξ is seperated from 0 in Dt by ρ. The next lemma is Lemma 2.1 in [23]: 20 ˜ be simply connected domains in C. Let ρ either be a Jordan curve Lemma 2. Let D ⊂ D ˜ which intersects ∂D or a crosscut in D. ˜ Let Z1 , Z2 be two connected subsets or prime in D ˜ such that D(ρ; ˜ Zj ) is well defined for both j = 1, 2, and are nonequal. This means ends of D ˜ ˜ that D\ρ is a neighborhood of both Z1 and Z2 in D, and Z1 is disconnected from Z2 in D by ρ. ˜ Let Λ be the set of connected Suppose that D is a neighborhood of Z1 and Z2 in D. components of D ∩ ρ. Then there exists a unique λ1 ∈ Λ such that D(λ1 ; Z1 ) = D(λ1 ; Z2 ), and if λ ∈ Λ such that D(λ; Z1 ) = D(λ; Z2 ), then D(λ1 ; Z1 ) ⊂ D(λ; Z1 ) and D(λ; Z2 ) ⊂ D(λ1 ; Z2 ). The λ1 obtained in Lemma 2 will be referred to as the first subcrosscut of ρ to seperate Z1 and Z2 in D. The conclusion of the lemma states that of all subcrosscuts of ρ in D which disconnect Z1 and Z2 , λ1 is closest to Z1 in the sense that the component containing Z1 it determines is contained in the component determined by any other such subcrosscut. 2.2.2 The covering space H∗ for D The upper half-plane can be seen as a covering space for the unit disc under the exponential map. Let H∗ be the cylinder defined by H∗ = {[z]∼ : z ∈ H}, where ∼ is the equivalence relation z ∼ w if z − w ∈ 2πZ. Under this equivalence relation, the map ei : H∗ → D\{0} defined by ei (z) = eiz is a conformal map. The boundary of H∗ is R∗ = R with the same equivalence relation. We can treat 0 as the preimage of ∞ under ei , and ei0 = 1. The map 1 i log : D\{0} → H∗ is also seen as a conformal map. Any set K ⊂ H∗ can be viewed as a 2π-periodic set K + 2πZ ⊂ H. An H∗ -hull is a set K ⊂ H∗ so that ei (K ) ⊂ D is a D-hull, and an H∗ -domain is a set H ⊂ H∗ so that 21 ei (H) ⊂ D is a D-domain. Therefore, an H∗ -domain is H∗ \K for an H∗ -domain K . Radial SLEκ can be studied by looking at the covering space H∗ . Let λ ∈ C[0, ∞) and z ∈ H∗ . The covering radial Loewner equation driven by λ is given by ∂t ht (z) = cot2 (ht (z) − λt ), h0 (z) = z, (2.3) where cot2 (z) = cot(z/2). Note that if (gt ) is the radial Loewner evolution driven by λ, then ht (z) = −i log gt (eiz ) satisfies (2.3). If λt = √ κBt , where Bt is a standard one dimensional Brownian motion, let γ be the preimage under ei of the radial SLEκ trace in D from 1 to 0 driven by λ. Then we call γ a radial SLEκ trace in H∗ from 0 to ∞, which is a 2π-periodic family of paths in the cylinder H∗ . If H is an H∗ -domain with a prime end w0 , we can define a radial SLEκ process in H from w0 to ∞ by mapping conformally from H to D = ei (H), to D, then back to H∗ with 1i log. To work in the cylinder H∗ , we must carefully define what distance and conformal radius ∗ mean in this domain. For z , w ∈ H which can be represented as z + 2πZ, w + 2πZ ∗ respectively for z, w ∈ H, then the distance from z to w in H is defined to be Euclidean distance between the sets z + 2πZ and w + 2πZ in H. It will be written as |z − w |∗ to distinguish from the distance between points in C. Then |z − w |∗ is the distance between closest representatives of equivalence classes. ∗ Similarly, if A , B ⊂ H with A = A + 2πZ and B = B + 2πZ, then the distance from A to B in H∗ is the Euclidean distance from A + 2πZ to B + 2πZ, and is denoted by ∗ distH∗ (A , B ). Given any z ∈ H with z = z + 2πZ, the ball of radius r centered at z is denoted by B(z , r), and is represented in H by B(z, r) ∩ H + 2πZ. Note that for r < π, then the representatives of B(z , r) are nonoverlapping in H. 22 In [2], several estimates are proven about one-point estimates with conformal radius instead of distance, so we need to provide the definition of conformal radius in H∗ . Given any simply connected domain D ⊂ C, and a point z ∈ D, the conformal radius of z in D is defined by radD (z) = 1/|φ (z)|, where φ : D → D is conformal with φ(z) = 0. It follows readily that radD (z) is a conformal invariant. That is, if φ : D → D is a conformal map and z ∈ D, then radφ(D) (φ(z)) = |φ (z)|radD (z). Since φ : H∗ → D\{0} defined by φ(z ) = eiz is a conformal map in the covering space, we can add ∞ into H∗ , thus adding 0 back into the image of φ : H∗ ∪ {∞} → D so that φ is conformal and the cylinder is simply connected. We can then use the conformal invarience formula to define conformal radius in the covering space H∗ , and any other H∗ -domain. If a set D is not connected, we define ΥD (z) to be the conformal radius of z in the component of D in which z lies. This will be relevant when z becomes cut off from the target of the SLE path by the curve. In this case, even though we define Dt to denote the component of D\γ[0, t] containing the target by time t, we will still use the notation ΥDt (z). For slightly easier calculations, we introduce the scaling ΥD (z) = radD (z)/2. From the definition, a quick calculation verifies that radD (z) = 1 − |z|2 , and so ΥD (z) = 1 − |z|2 . For 2 any z ∈ H∗ , we can show that ΥH∗ (z ) = sinh(y ), (2.4) where y = Im(z ). To see this, note that the conformal invarience definition shows that radH∗ (z ) = 1 |ieiz | (1 − |eiz |2 ) = |e−iz | − |eiz | = ey − e−y , and dividing by 2 gives (2.4). Remark 1: We will need to use applications of Koebe’s distortion theorem for the map 23 ei on sets in H∗ . For this to be valid, the set will have to sit inside B(z, r) + 2π for some r < π so that ei can be restricted to a conformal map on a subdomain of C. Remark 2: In [2], the covering space H∗ is defined slightly differently than we do by assuming z ∼ w if z−w ∈ πZ. In this case, the associated conformal map is φ(z ) = e2iz , and the same calculation as above gives radH∗ (z ) = 2 cosh(z ) sinh(z ) = sinh(2y ). However, this slight change in coordinates only affects the estimates by a constant factor. 2.2.3 Extremal length and conformal transformations Let dΩ (A, B) denote the extremal distance from A to B in Ω. For the definition of extremal distance, see [1]. Note that this is distinct from the notations dist(A, B) and distH∗ (A, B), both of which represent Euclidean distance. Define Λ(R) = dΩ ([−1, 0], [R, ∞)), where R ΩR = C\{[−1, 0] ∪ [R, ∞)}. By Teichm¨ uller’s theorem [1], this is maximal among doubly connected domains in modulus (extrememal distance between boundary components) which separate {−1, 0} and {w, ∞} with |w| = R. Moreover, Λ(R) ≤ 1 log(16(R + 1)) for all 2π values of R ≥ 1. We are going to need to prove some estimates about extremal length which are conformally invariant, and so are more flexible than estimates involving Euclidean distance. Moreover, we are going to look at extremal distance in the cylinder H∗ . Since H∗ looks locally like H, if two sets are sufficiently close to each other, the extremal distance between them should behave the same as the extremal distance in H. Lemma 3. Let η be a crosscut in {z ∈ H : Re(z) > 0} with endpoints 0 < a < b. Define r = sup{|z − a| : z ∈ η}. Then there is a constant C < ∞ so that, if r < a, r ≤ Ce−πdH (η,(−∞,0]) . a 24 (2.5) Proof. Let Ω be the component of H\η whose boundary contains 0, and define Ω = Ω ∪ (0, a) ∪ (b, ∞). Let Ωdoub = Ω ∪ {z : z ∈ Ω}, and let η doub be the component of the boundary of Ωdoub which does not contain 0, which is the closure of the double of η. Then dH ((−∞, 0], η) = 2dΩdoub ((−∞, 0], η doub ). Fix some w in the closure of η with |w − a| = r, and let ξ be the argument of w − a. Then φ(z) = (z − a)e−iξ /r − 1 is the affine transformation with φ(a) = −1, φ(w) = 0, and ˆ = φ(Ωdoub ), then Ω ˆ is a doubly connected domain seperating |φ(0)| = |ae−iξ /r + 1|. If Ω {−1, 0} from {φ(0), ∞}. By the conformal invarience of extremal distance and Teichm¨ uller’s theorem, we have doub ))) ≤ dΩdoub ((−∞, 0], η doub ) = dΩ ˆ (φ((−∞, 0], φ(η ≤Λ 1 Λ |ae−iξ /r + 1| 2π a 1 1 a +1 ≤ ln(16(a/r + 1) + 1) ≤ ln 33 . r 2π 2π r Rearranging this proves (2.5) for r small. The following application of Koebe’s distortion theorem will be used repeatedly in the next section to show that interior estimates are comparable after applying conformal maps. We also derive a growth estimate for the inverse of the covering map for the cylinder. Lemma 4. a) Let D ⊂ C be a domain, and assume that B(z0 , R) ⊂ D. Suppose φ is a conformal map defined on D, and let M =dist(z0 , ∂D). Suppose that r < R/7, and define ˜= R R|φ (z0 )| r|φ (z0 )| , r˜ = . 2 (1 + R/M ) (1 − r/M )2 25 Then ˜ ⊂ φ(B(z0 , R)), φ(B(z0 , r)) ⊂ B(φ(z0 ), r˜) ⊂ B(φ(z0 ), R) and there is an absolute constant C < ∞ so that r r˜ r ≤ ≤C . ˜ R R R b) Let D be a D-domain, and let H be an H∗ -domain with ei (H) = D. Let z0 ∈ D and z0 ∈ H so that eiz0 = z0 . Then there is a constant C < ∞ so that if y0 = 1−|z0 | ≤ 1/2, we have dist(z0 , ∂D) ≤ CdistH∗ (z0 , ∂H). Proof. Applying the Growth Theorem to the univalent map f : D → C defined by f (w) = φ(M w + z0 ) − φ(z0 ) , M φ (z0 ) we get that if |z − z0 | = ρ ∈ (0, M ), then |φ(z) − φ(z0 )| ρ/M ρ/M ≤ ≤ . M |φ (z0 )| (1 + ρ/M )2 (1 − ρ/M )2 (2.6) ˜ is M |φ (z0 )| times Then r˜ is M |φ (z0 )| times the right hand side of (2.6) for ρ = r, and R the left hand side for ρ = R. Then if r (1 + R/M )2 r˜ = ˜ R (1 − r/M )2 R 26 is smaller than 1, we get the desired containments. If r < R/7 and R < M , then r˜ 1 4 ≤ < 1. ˜ 7 (6/7)2 R Moreover, 1≤ 4 (1 + R/M )2 ≤ , (1 − r/M )2 (6/7)2 which proves the comparability. To prove part b) let, φ(z) = 1i log(z) be the map from D\{0} to H. Let r =dist(z0 , ∂D), so that the assumption y0 ≤ 1/2 implies that a single valued branch of φ is conformal on the open ball B(z0 , r). By Koebe’s 1/4 theorem, B(z0 , r/4) ⊂ φ(B(z0 , r)) ⊂ H. If w ∈ ∂H, this implies that |z0 − w |∗ ≥ r/4 = dist(z0 , ∂D)/4. 2.3 Interior and boundary estimates The following estimate is Proposition 5.1 in [2]. Proposition 1. (Interior estimate 1) Let γ be a radial SLEκ curve in H∗ from 0 to ∞. There exists constants 0 < c1 < c2 < ∞ so that if z ∈ H∗ with y = Im(z ) ≤ 1 and ≤ 1/2, then c1 y |z |∗ α 2−d ≤ P Υ∞ (z ) ≤ Υ0 (z ) ≤ c2 where Υt (z ) is the conformal radius of z in H∗ \γ [0, t]. 27 y |z |∗ α 2−d , Lemma 5. (Interior estimate 2) Let γ be a radial SLEκ trace in H∗ from 0 to ∞. Then there is a constant C < ∞ so that if z ∈ H∗ with y =Im(z ) ≤ ln(2) and r < y , P[distH∗ (z , γ ) ≤ r] ≤ C y |z |∗ α r y 2−d =C Py (r) Py (|z |∗ ) . Remark: By restricting to y ≤ ln(2), we are restricting slightly further than we are in Proposition 1. We do this restriction so that we can use Lemma 4, part b). Later, we will use part a) of Lemma 4 applied to a Loewner map to extend to the whole space. Proof. Let z ∈ D so that eiz = z, denote the H∗ -hull determined by γ [0, t] by Ht , and let Dt = ei (Ht ) for each t. By the definition of conformal radius of domains in H∗ , we have ΥDt (z) = |(ei ) (z )|Υt (z ) = e−y Υt (z ) for all t < ∞. By Koebe’s 1/4 theorem and the assumption y ≤ log(2), we have Υt (z ) = ey ΥDt (z) ≤ Cdist(z, ∂Dt ). Part b) of Lemma 4 implies that dist(z, ∂Dt ) ≤ CdistH∗ (z , ∂Ht ), and therefore Υt (z ) ≤ CdistH∗ (z , ∂Ht ) ≤ CdistH∗ (z, γ [0, t]). Taking t → ∞ gives Υ∞ (z ) ≤ cdistH∗ (z , γ ) for an absolute constant c < ∞. By the above paragraph, for small , we have {distH∗ (z , γ ) < Υ0 (z ) } ⊂ {Υ∞ (z ) ≤ c Υ0 (z )}. Therefore, P[distH∗ (z , γ ) ≤ Υ0 (z ) ] ≤ C c 28 y |z |∗ α 2−d . Renaming variables and picking sufficiently small depending on z and r, we can let r = Υ0 (z ) . Then c P[distH∗ (z , γ ) ≤ r] ≤ C y |z |∗ α 2−d cr ≤C Υ0 (z ) y |z |∗ α r y 2−d , where the last inequality follows from (2.4), which says that Υ0 (z ) = ΥH∗ (z ) = sinh(y ) ≥ y . This is equal to Py (r)/Py (|z |∗ ), since |z |∗ ≥ y implies that Py (|z |∗ ) = |z |α ∗ and r < y implies that Py (r) = (y )α−(2−d) r2−d . The initial boundary estimate we will work with is Lemma 5.1 in [2]. Lemma 6. (Boundary estimate 0) There is a constant C < ∞ so that if γ is a radial SLEκ curve in H∗ , x ∈ R∗ , and r < |x |∗ , then P[dist(x , γ ) ≤ r] ≤ C α r . |x |∗ We want to modify this estimate into a more general and conformally invariant version which can be applied in more general domains. In the next lemma, we will derive an estimate involving the extremal distance between two crosscuts in the cylinder H∗ . This will be determined by the domain between the crosscuts, and will be the same as the extremal distance between a representation of each of them in H. Lemma 7. Let D be any D domain, and let γ be radial SLEκ in D from a prime end w0 of D to 0. Let ρ, η be crosscuts in D with small radius, η contained in the component of D\ρ 29 distinct from the component containing 0 and w0 . Then P[γ ∩ η = ∅] ≤ Ce−απdD (ρ,η) . (2.7) Similarly, the inequality (2.7) holds when γ is radial SLEκ in an H∗ domain D from w0 to ∞, and ρ, η are non-self-intersecting 2π-periodic crosscuts in H∗ so that ρ separates w0 and ∞ from η. Proof. In either case of the lemma, let φ be a conformal map from the domain to H∗ sending w0 to 0. Then ρ = φ(ρ) is a non-self-intersecting crosscut in H∗ which separates 0, ∞ from η = φ(η), and φ(γ) = γ˜ is a time changed radial SLEκ curve in H∗ . Let η0 be the representative for η in H with η0 ⊂ {z ∈ H : 0 < Re(z) ≤ 2π}, and define ρ0 similarly. Denote the endpoints of η0 by 0 < a < b < 2π, and let x = min{a, 2π − b}. If x = a, define r = sup{|z − a| : z ∈ η0 }. If x = 2π − b, let r = sup{|z − b| : z ∈ η0 }. Either way, η0 ⊂ B(x, r), and so by Lemma 6, we have P[γ ∩ η = ∅] = P[˜ γ ∩ η = ∅] ≤ P[dist(˜ γ , x) ≤ r] ≤ C r α . x Applying Lemma 3 to η0 , we have r ≤ C e−πdH (η0 ,(−∞,0]) ∨ e−πdH (η0 ,[2π,∞)) ≤ Ce−πdH (ρ0 ,η0 ) , x where the last inequality is due to the comparison principle of extremal length. Combining these two inequalities gives P[γ ∩ η = ∅] ≤ Ce−απdH (ρ0 ,η0 ) = Ce−απdD (ρ,η) , 30 where the last equality is from the conformal invariance of extremal length, and since it is the domain between the crosscuts which determines the length. Now we can combine the interior and boundary estimates into the general one-point estimate for the cylinder. We now also remove the assumption that y0 ≤ ln(2) to extend the estimate to the whole space. Lemma 8 (One-point estimate for cylinder). Suppose that H is an H∗ domain, and γ is radial SLEκ in H from a prime end w0 to ∞. Fix z0 ∈ H∗ with imaginary part y0 , and let π > R > r > 0. Define ρ = {z ∈ H∗ : |z − z0 |∗ = R}, η = {z ∈ H∗ : |z − z0 |∗ = r}. Moreover, suppose that {z ∈ H∗ : |z − z0 |∗ ≤ R} ⊂ D and that w0 ∈ / {x ∈ R∗ : |z0 − x |∗ < R}. Then P[γ ∩ η = ∅] ≤ C Py (r) 0 Py (R) . 0 Proof. Observe that, by assumption, the representatives for ρ in H under the equivalence relation consists of multiple disconnected components, rather than having overlap. This means any of the covering maps are locally conformal at each piece of ρ, and everything inside of ρ, so Lemma 4 will be applicable. The proof breaks down into three cases, each depending on how far from R∗ is the point z0 . The first case is the far away case, when y0 > R. Also, we must first assume that y0 ≤ ln(2). Let hH : H → H∗ be the canonical conformal map taking w0 to 0 and ∞ to ∞, so that γ˜ = hH (γ ) is a time-changed radial SLEκ curve from 0 to ∞ . Let z˜0 = hH (z0 ), and let y˜0 = Im(˜ z0 ). Since hH is conformal on a simply connected neighborhood of any component of ρ, Lemma 4 a) implies that, assuming r < R/7 without loss of generality (else 31 we modify the constant by a factor of 7), z0 , γ˜ ) ≤ r˜] P[γ ∩ η = ∅] = P[˜ γ ∩ hH (η) = ∅] ≤ P[distH∗ (˜ Py˜0 (˜ r) Py˜ (˜ r) ≤C ≤C 0 =C ˜ Py˜0 (|˜ z0 |∗ ) Py˜0 (R) r˜ 2−d r 2−d ≤C , ˜ R R ˜ are defined as in Lemma 4 with respect to the conformal map r, R, and hH . where r˜, R Assume now that z0 ∈ H is arbitrary, and let ht : H\γ [0, t] → H be the Loewner maps for the radial SLEκ process in H. For any T > 0, define γ T (t) = γ (T + t), so that conditioned on FT , γ T is a radial SLEκ trace from γ (T ) to ∞. Define zt = ht (z0 ), and define a stopping time by τ = inf{t ≥ 0 : Im(zt ) ≤ ln(2)}. Define rτ , Rτ as in Lemma 4 at z0 with respect to r, R, hτ . Then Lemma 4 implies that P[distH∗ (zτ , γτ ) < r|Fτ , τ < ∞] ≤ C rτ 2−d r 2−d ≤C . Rτ R Since P[τ < ∞] ≤ 1, this proves the lemma for all z0 ∈ H. For case 2, we consider the close case, 0 < y0 < r. We will use the boundary estimates r α to derive an upper bound of . By modifying the constant slightly, we can assume that R R > 4r. Then in order to cross from ρ to η, γ must also pass through ρ = {z ∈ H∗ : |z − Re(z0 )|∗ = R/2} and η = {z ∈ H∗ : |z − Re(z0 )|∗ = 2r}, which are two semicircles such that dD (ρ , η ) = (1/π) log(R/4r). Using Lemma 7, we get that P[γ ∩ η = ∅] ≤ P[γ ∩ η = ∅] ≤ Ce−απdD (η ,ρ ) = C r α . R Lastly, we have to mix interior and boundary estimates in the middle distance case when 32 0 < r < y0 < R. Let ρ = {z ∈ H∗ : |z − z0 |∗ = y0 } which is a circle tangent to R∗ between η and ρ. Let T = inf{t > 0 : γ (t) ∈ ρ }, which is a stopping time so that {γ ∩ η = ∅} ⊂ {T < ∞}. Using case 2, we can see that P[T < ∞] ≤ C Py (y0 ) 0 Py (R) . 0 Define γ T (t) = γ (T + t). By the domain Markov property and the first case, we can see that P[γ T ∩ η = ∅|γ [0, T ], T < ∞] ≤ C Py (r) 0 Py (y0 ) 0 . Combining these two inequalites gives case 3. Using the Growth Theorem estimate again, we can prove the analogous estimate for radial SLE in the disc. Lemma 9 (One-point estimate for disc). Let D be a D domain with a prime end w0 , and let γ be a radial SLEκ curve in D from w0 to 0. Let z0 ∈ D, y0 = 1 − |z0 |. Suppose that 0 < r < R < dist(z0 , {0, w0 }). Let ρ = {z ∈ D : |z − z0 | = R}, and η = {z ∈ D : |z − z0 | = r}, and assume that {z ∈ D : |z − z0 | ≤ R} ⊂ D. Then P[γ ∩ η = ∅] ≤ C Py0 (r) . Py0 (R) Proof. Let φ : D\{0} → H∗ be defined by φ(z) = 1i log(z), and suppose z0 ∈ H := φ(D) so that eiz0 = z0 . The proof will break into the same 3 cases as in the proof of Lemma 8: 33 R < y0 , r > y0 , and r < y0 < R. In the first two cases, we want to prove that P[dist(z0 , γ)] ≤ C r p R for p ∈ {α, 2 − d}, so it will suffice to prove an upper bound for r/R which will work for both cases simultaneously. The third case can be handed similarly to the third case in Lemma 8. By assumption, φ is a conformal map on a domain which has positive distance from ˜ as in Lemma 4 B(z0 , R). Let γ be a radial SLEκ trace in H from φ(w0 ) to ∞. Define r˜, R with respect to φ and z0 . If r < R/7, the inclusions in Lemma 4, part a), imply that in the ˜ and in the boundary interior case, we can apply the interior estimate of Lemma 8 to z0 , r˜, R, case we can apply the boundary estimate. In either case, P[dist(z0 , γ) ≤ r] ≤ P[distH∗ (z0 , γ ) ≤ r˜] ≤ C r˜ p r p , ≤C ˜ R R where p ∈ {α, 2 − d} is chosen appropriately. Lemma 10. Let D be a D domain with a prime end w0 , and let γ be a radial SLEκ curve in D from w0 to 0. Let ρ be a crosscut in D such that D∗ (ρ) is not a neighborhood of w0 ˜ be a domain that contains D, and ρ˜ be a subset of D ˜ that in D, and let S ⊂ D∗ (ρ). Let D ˜ which intersects ∂D or a crosscut in D. ˜ contains ρ. Let η˜ be either a Jordan curve in D ˜ Then Suppose that η˜ disconnects ρ˜ from S in D. −απd ˜ (˜ ρ,˜ η) D . P[γ ∩ S = ∅] ≤ Ce Proof. By Lemma 2, η˜ has a subcrosscut η in D which disconnects S from ρ. Since S ⊂ 34 D∗ (ρ), we have η ⊂ D∗ (ρ) and S ⊂ D∗ (η). Therefore, D(ρ; η) = D∗ (ρ) is neither a neighborhood of 0 nor w0 in D. Using the boundary estimate from Lemma 7, we see that P[γ ∩ S = ∅] ≤ P[γ ∩ D∗ (η) = ∅] ≤ Ce−απdD (ρ,η) ≤ C −απd ˜ (˜ ρ,˜ η) D . The last inequality follows from the comparison principle for extremal length, since any path ˜ which connects ρ˜ and η˜. in D connecting ρ and η is a subarc of itself contained in D 2.4 Components of crosscuts Before we state the main theorem of this section, we will introduce the notation to be used. Let Ft be the right continuous filtration determined by the radial SLE curve γ. For any set S ⊂ D, let τS = inf{t ≥ 0 : γ(t) ∈ S}. For any stopping time τ , define γ τ (t) = γ(τ + t). Define Dt = D(γ[0, t]). Theorem 4. Let γ be a radial SLEκ curve in D from 1 to 0. Suppose that z0 , z1 , . . . , zm ∈ D\{0, 1}. For each zj , let 0 < rj ≤ Rj , and define the circles ξˆj = {|z − zj | = Rj } and ξj = {|z − zj | = rj }. Assume that neither 0 nor 1 are contained in D∗ (ξˆj ; zj ) for each j, and that D∗ (ξˆj ) ∩ D∗ (ξˆi ) = ∅ for j = j. Let r0 ∈ (0, r0 ) and define ξ0 = {|z − z0 | = r0 }. Define the event E = {τξ0 < τξˆ ≤ τξ1 < · · · < τξˆ ≤ τξm < τξ < ∞}. m 1 0 35 If yj = 1 − |zj |, then P[E|Fτξ ] ≤ C m 0 m P (r ) yj j r0 α/4 . R0 Pyj (Rj ) j=1 Note that the assumptions here imply that neither the start point, 1, nor the endpoint 0, of the SLE curve are enclosed by the discs ξˆj . We also assume interiors of circles with different centers have no overlap, and that the boundary circles do not meet. The proof is similar to the proof of Theorem 3.1 in [23], but for completeness we include complete details. Proof. Consider the discs ξ which intersect the boundary. We know that the probability that γ hits the points in ξ ∩ ∂D is equal to 0, and so τξ = τξ∩D a.s. Therefore, we can assume that each ξ is either a Jordan curve or a crosscut in D. For each j = 0, 1, . . . , m, let τj = τξj and τˆj = τξˆ , and define τm+1 = τξ . j 0 By the Domain Markov Property of SLE and Lemma 9, we see that P[τj < ∞|Fτˆj ] ≤ C Pyj (rj ) Pyj (Rj ) . (2.8) Combining these together gives m P[E|Fτ0 ] ≤ Cm j=1 Pyj (rj ) Pyj (Rj ) . If r0 = R0 , then we are done. Suppose that R0 > r0 . Create a new arc ρ = {z ∈ D : |z − z0 | = √ R0 r0 }, so that ρ is either a Jordan curve or a crosscut in D between ξ0 and ξˆ0 . Therefore, we know that log(R0 /r0 ) dD (ρ, ξ0 ), dD (ρ, ξˆ0 ) ≥ . 4π 36 (2.9) Note that ρ seperates ξ0 from 0. In the following argument, we will need to keep track of how the D-domains Dt are divided by ρ at any particular time. Let T = inf{t ≥ 0 : ξ0 ⊂ Dt }. Then if τ0 ≤ t < T , ξ0 is a connected subset of Dt . In this case, since the starting point 1 is outside of ξˆ0 and γ intesercts ξ0 , it must be that γ intersects ρ, and so ρ intersects ∂Dt . By Lemma 2, there is a first subcrosscut of ρ in Dt , to be denoted ρt , which seperates ξ0 from 0 for each τ0 ≤ t < T . Now, we need to break the event E into several cases based on the behavior of the curve γ as it intersects the circles in the correct order. Let I = {(j, j + 1) : 0 ≤ j ≤ m} ∪ {(j, j) : 1 ≤ j ≤ m}, and define a sequence of events {Ai : i ∈ I} by 1. A(0,1) = {T > τ0 } ∩ {D∗ (ξ1 ) ⊂ Dτ∗0 (ρτ0 )} 2. A(j,j) = {T > τj } ∩ {D∗ (ξj ) ⊂ Dτj−1 (ρτj−1 )} ∩ {D∗ (ξj ) ⊂ Dτ∗ (ρτj )}, 1 ≤ j ≤ m. j 3. A(j,j+1) = {T > τj } ∩ {D∗ (ξj ) ⊂ Dτj (ρτj )} ∩ {D∗ (ξj+1 ) ⊂ Dτ∗ (ρτj )}, 1 ≤ j ≤ m − 1. j 4. A(m,m+1) = {T > τm } ∩ {D∗ (ξm ) ⊂ Dτm (ρτm )}. Observe that for each j, the events A(j,j) , A(j,j+1) are Fτj measurable. What I claim is that E ⊂ ∪i∈I Ai . (2.10) To see this, observe that A(m,m+1) is the event that at time τm , ξm lies outside of ρ, relative to 0 and the path γ. If that does not happen, then ξm must lie inside ρ at time τm . Now suppose that A(m,m) does not happen, which is the event that at time τm−1 , ξm lies outside of ρ, but at time τm , ξm lies inside of ρ. Then it must be that ξm lies inside of ρ at τm−1 . Proceeding along this way inductively proves (2.10). Complete details can be found 37 in Section 2.7. Now it will suffice to show that P[E ∩ Ai |Fτ0 ] ≤ Cm m P (r ) yj j r0 α/4 R0 Pyj (Rj ) (2.11) j=1 for each i ∈ I. We will break it down to the four cases i = (m, m + 1), (j, j), (j, j + 1), and (0, 1). In all of these cases, we will use the convention γ T (t) = γ(T + t) for any T, t ≥ 0. Case (0, 1). Suppose that A(0,1) occurs, and that τ0 < τˆ1 . First, we claim that ξˆ1 ⊂ D∗τ0 (ρτ0 ). Note that both ξˆ1 and D∗ (ξ1 ) are contained in D∗ (ξˆ1 ) ∪ ξˆ1 , which is a connected subset of the disjoint union Dτ0 (ρτ0 ) ∪ Dτ∗0 (ρτ0 ). By assuming that the event A(0,1) occurs, however, we can conclude that they must both be contained in Dτ∗0 (ρτ0 ), which proves the claim. Also, note that ρ disconnects ξˆ1 from ξ0 in D, and ρ must intersect ∂Dτ0 . By Lemma 2, there is a subcrosscut ρτ0 of ρ which is first to seperate ξˆ1 from ξ0 in the domain Dτ0 . Since both ξˆ1 and ξ0 lie in Dτ∗0 (ρτ0 ), so does ρτ0 . Note that this implies that ρτ0 = ρτ0 , and that Dτ∗0 (ρτ0 ) ⊂ Dτ∗0 (ρτ0 ). Since ρτ0 was defined to be the first subcrosscut of ρ in Dτ0 that disconnects ξ0 from 0, and since Dτ∗0 (ρτ0 ) is contained in the domain determined by ρτ0 , it cannot be that ρτ0 disconnects ξ0 from 0. Therefore, we conclude that ξ0 ⊂ Dτ0 (ρτ0 ) and ξˆ1 ⊂ Dτ∗0 (ρτ0 ). Observe that D∗ (ξ0 ) is a connected subset of Dτ0 \ρτ0 , and contains ξ0 and a curve which approaches γ(τ0 ) ∈ ξ0 . Therefore, Dτ0 (ρτ0 ; γ(τ0 )) = Dτ0 (ρτ0 ; ξ0 ) = Dτ0 (ρτ0 ). It follows that Dτ0 (ρτ0 ; ξˆ1 ) = Dτ∗0 (ρτ0 ) is not a neighborhood of γ(τ0 ) = γ τ0 (0), where γ τ0 (conditioned on Fτ0 ) is a radial SLEκ curve in the D-domain Dτ0 . Since τ0 < τˆ1 , the event 38 {ˆ τ1 < ∞} implies that the conditioned SLE curve γ τ0 visits ξˆ1 . Since ξˆ0 disconnects ξˆ1 from ρτ0 ⊂ ρ in D, and ξˆ0 intersects ∂Dτ0 , we can apply Lemma 10 to conclude that ˆ P[ˆ τ1 < ∞|Fτ0 , A(0,1) , τ0 < τˆ1 ] ≤ Ce−αdD (ρ,ξ0 ) ≤ C r0 α/4 . R0 Note that the second inequality follows from (2.9). Combining the above inequality with inequality (2.4) proves that inequality (2.11) holds for the case i = (0, 1). Case (j, j + 1), for 1 ≤ j ≤ m − 1. Suppose that A(j,j+1) occurs, and τj < τˆj+1 . By the same argument used in the case for A(0,1) , we can conclude that there exists a subcrosscut of ρ, which we will call ρτ , which disconnects ξˆj+1 from ξ0 in Dτj . It follows that Dτ∗ (ρτ ) ⊂ j j j Dτ∗ (ρτj ), and then we can conclude that ξ0 ⊂ Dτj (ρτ ) and ξˆj+1 ⊂ Dτ∗ (ρτ ). Since D∗ (ξj ) j j j j is a connected subset of Dτj (ρτj ) and contains a curve approaching γ(τj ) ∈ ξj , we can see that Dτj (ρτj ; γ(τj )) = Dτj (ρτj ; D∗ (ξj )) = Dτj (ρτj ). Therefore, Dτ∗ (ρτ ) ⊂ Dτ∗ (ρτj ) is not j j j τ a neighborhood of γ(τj ) = γ j (0). τ Since τj < τˆj+1 , the event {ˆ τj+1 < ∞} implies that γ j , the radial SLEκ curve in Dτj after conditioning on Fτj , visits ξˆj+1 . Since ξˆ0 disconnects ξˆj+1 from ρ, and therefore from ρτ , in D, Lemma 10 implies that j ˆ P[ˆ τj+1 < ∞|Fτj , A(j,j+1) , τj < τˆj+1 ] ≤ Ce−απdD (ρ,ξ0 ) ≤ C r0 α/4 . R0 As in the above case, this proves inequality (2.11) for the case i = (j, j + 1) when 1 ≤ j ≤ m − 1. Case (m, m + 1). Suppose that τm < τm+1 , and that A(m,m+1) occurs. The set D∗ (ξm ) is connected and contained in Dτm \ρτm , and also γ(τm ) ∈ ξm . This implies that Dτm (ρτm ; γ(τm )) = 39 Dτm (ρτm ; D∗ (ξm )) = Dτm (ρτm ). Therefore, Dτ∗m (ρτm ) is not a neighborhood of γ τm (0) = γ(τm ) in Dτm . Since we are assuming that τm < τm+1 , the event {τm+1 < ∞} implies that the curve γ τm , which after conditioning on Fτm is a radial SLEκ in Dτm from γ(τm ) to 0, must visit ξ0 ⊂ Dτ∗m (ρτm ). Since ξ0 disconnects ξ0 from ρ in D and intersects ∂Dτm , we can apply Lemma 10 to show that P[τm+1 < ∞|Fτm , A(m,m+1) , τm < τm+1 ] ≤ Ce−απdD (ξ0 ,ρ) ≤ C r0 α/4 . R0 Case (j, j), for 1 ≤ j ≤ m. We now prove inequality (2.11) for A(j,j) . Fix a j in {1, . . . , m}, and define σj = inf{t ≥ τj−1 : D∗ (ξj ) ⊂ Dt∗ (ρt )}. This can be seen as the first time after τj−1 that the SLE curve γ hits the crosscut ρ on the “correct” side of ξˆj , i.e. γ cuts the disc off from 0. There are a few observations to make about σj which follow from Lemma 11 (in Section 2.7). First, σj is an Ft -stopping time. If σj < ∞, then D∗ (ξj ) ⊂ Dσ∗ (ρσj ). j If the event A(j,j) occurs, then so does {τj−1 < σj < τj }. Finally, if τj−1 < σj < ∞, then it must be that γ(σj ) is an endpoint of ρσj . This implies that Dσ∗ (ρσj ) is neither a j neighborhood of γ(σj ) nor of 0. We define two events which seperate the event A(j,j) . Define F< = {σj < τˆj }, and F≥ = {τj > σj ≥ τˆj }. Notice that A(j,j) ⊂ F< ∪ F≥ , so if we can prove (2.11) for both F< and F≥ instead of A(j,j) , the case will be proven. First, assume that F< happens. Then D∗ (ξˆj )∪ ξˆj is a connected subset of (D\γ[0, ρσj ])\ρ that contains D∗ (ξj ), and so ξˆj ⊂ Dσ∗ (ρσj ; D∗ (ξj )) = Dσ∗ (ρσj ). Since ξˆ0 disconnects ρ from j j 40 ξˆj in D, Lemma 10 and (2.9) imply that ˆ P[ˆ τj < ∞|Fσj , F< ] ≤ Ce−απdD (ρ,ξ0 ) ≤ C r0 α/4 . R0 This implies that P[τj < ∞, F< |Fτj−1 ] ≤ C r0 R) α/4 Pyj (rj ) Pyj (Rj )) . Next, we assume that F≥ occurs, which is the more difficult case. Define N = log(Rj /rj ) , where · is the integer valued ceiling function. In the event F≥ , the SLE curve passes through the outer circle ξˆj , then hits the crosscut ρ before returning to hit the inner circle ξj . What we need to do is divide the annulus {rj ≤ |z − zj | < Rj } into N subannuli until we can identify the last subcircle to be crossed before the time σj . j 1/N Define the circles ζk = {|z − zj | = RjN −k rj } for 0 ≤ k ≤ N . Note that ζ0 = ξˆj , ζN = ξj , and a higher index k indicates that ζk is more deeply nested inside ξˆj . Then F≥ ⊂ ∪N k=1 Fk , where Fk = {τζ k−1 ≤ σj < τξ }. ζ k If the event Fk occurs, then ζk ⊂ Dσ∗ (ρσj ). This is because D∗ (ζk )∪ζk is a connected subset j of D\γ[0, σj ] \ρ which contains ζk , D∗ (ξj ), and Dσ∗ (ρσj ). By Lemma 10 and inequality j (2.9), we conclude that ˆ P[τζ < ∞|Fσj , Fk ] ≤ Ce−απdD (ρ,ζk−1 ) ≤ Ce−απ(dD (ρ,ξ0 )+dD (ζ0 ,ζk−1 )) k ≤C r0 α/4 R0 rj α(k−1))/2N . Rj 41 (2.12) By Lemma 9, we get that P[Fk |Fτj−1 , τj−1 < τˆj ] ≤ C P[τj < ∞|Fζ , Fk ] ≤ C k Pyj (RjN −k+1 rk−1 )1/N Pyj (Rj ) Pyj (rj ) Pyj (RjN −k rjk )1/N , . Combining the above three inequalities and the upper bound in Lemma 1 gives P[τj < ∞, Fk |Fτj−1 , τj−1 < τˆj ] ≤ C r0 α/4 R0 α(k−1) 2N rj Rj rj −α/N Pyj (rj ) . Rj Pyj (Rj ) Since F≥ ⊂ ∪N k=1 Fk , adding up the above inequality yields P[τj < ∞, F≥ |Fτj−1 , τj−1 < τˆj ] ≤ C r0 α/4 Pyj (rj ) R0 Pyj (Rj ) rj −α/N 1 − (rj /Rj )α/2 . Rj 1 − (rj /Rj )α/2N In Section 2.7, we prove that rj −α/N 1 − (rj /Rj )α/2 eα ≤ , Rj 1 − (rj /Rj )α/2N 1 − e−α/4 (2.13) and so we get P[τj < ∞, F≥ |Fτj−1 , τj−1 < τˆj ] ≤ C r0 α/4 Pyj (rj ) , R0 Pyj (Rj ) which is the same bound as that acheived for F< , though with a different constant. Com- 42 bining these two inequalities gives P[τj < ∞, A(j,j) |Fτj−1 , τj−1 < τˆj ] ≤ C r0 −α/4 Pyj (rj ) , R0 Pyj (Rj ) which completes the proof of (2.11) in the final case, and so the theorem follows. 2.5 Concentric circles Let Ξ be a family of mutually disjoint circles in C with centers in D\{0}, none of which pass through or enclose 0 or 1. We can define a partial order on Ξ by ξ1 < ξ2 if ξ2 is enclosed by ξ1 . Note that the larger circle has a smaller radius than the larger circle because the order is determined by visiting time. If ξ1 < ξ2 , any continuous path which hits both circles must pass through ξ1 before it hits ξ2 . When γ is an SLE curve in D starting from 0, ξ1 < ξ2 means that τξ1 < τξ2 , assuming γ passes through ξ2 . Also, observe that circles in Ξ are not necessarily contained in D. In fact, we want to account for circles in Ξ which have center in the boundary as well. Further, we assume that Ξ has a partition ∪e∈E Ξe with the following properties: 1. For each e ∈ E, the elements of Ξe are concentric circles whose radii form a geometric sequence with a common ratio of 1/4. For each e, let the common center be ze . Note that elements of Ξe are totally ordered. Let Re be the radius of the smallest circle (in the ordering on Ξ), and let re be the radius of the largest circle. Then there is some integer M ≥ 0 with Re = re 4M . 2. Let Ae = {z ∈ C : re ≤ |z − ze | ≤ Re } denote the closed annulus containing all of the circles in Ξe . Then we assume that the collection of annuli {Ae }e∈E is mutually 43 disjoint. We make a couple of quick remarks about the generality of this assumption. It may be that |Ξe | = 1, in which case re = Re , and the annulus Ae is just the single circle contained in Ξe . Also, if e1 = e2 , that does not necessarily mean that ze1 = ze2 . There can be multiple sets of concentric circles in Ξ with the same center with gaps between them. In this case, if members of Ξe2 have the smaller radii, we have Re2 < re1 . Also, there can be two components of the partition Ξe1 , Ξe2 so that each element in Ξe2 is ordered larger than each element of Ξe1 . In this case, the annulus Ae2 is contained in the bounded component of C\Ae1 . Theorem 5. Let Ξ be a family of circles with the properties listed above, and assume that γ is a radial SLEκ curve in D from 1 to 0. For each e ∈ E, let ye = 1 − |ze |. Then there exists a constant C|E| < ∞, which only depends on κ and the size of the partition |E|, so that P[∩ξ∈Ξ {γ ∩ ξ = ∅}] ≤ C|E| e∈E Pye (re ) . Pye (Re ) The strategy is to consider all possible orders σ that the SLE curve γ can visit all elements of Ξ. Under σ, γ may pass through several elements of a family Ξe0 , leave and visit other Ξe ’s, before returning to pass through the more inner circles in Ξe1 . Theorem 4 provides an estimate of the price paid by γ in order to return to the interior circles of Ξe0 . This gives an estimate for the probability of ∩ξ {γ ∩ ξ = ∅} in the prescribed order σ. We then add up over all appropriate orders σ and show that the constant only depends on |E|. Proof. Define S to be the set of permutations σ : {1, 2, . . . , |Ξ|} → Ξ such that ξ1 < ξ2 implies σ −1 (ξ1 ) < σ −1 (ξ2 ). Then S is the set of viable orders in which γ can visit the elements of Ξ for the first time. For an ordering σ ∈ S, σ(j) ∈ Ξ is the j-th circle visited by γ. Define the 44 event E σ = {τσ(1) < · · · < τσ(|Ξ|) < ∞}. Then E := ∩ξ∈Ξ {γ ∩ ξ = ∅} = ∪σ∈S E σ . What we need to do is bound P[E σ ]. Fix some σ ∈ S. Our first goal is to create a subpartition of {Ξe }e∈E into {Ξi }i∈I , where the elements of Ξi receive first visits from γ without interruption in the event E σ . e . Define For each e ∈ E, let Ne = |Ξe | − 1, and label the elements of Ξe by ξ0e < · · · < ξN e Je ⊂ {0, 1, . . . , Ne } by e ) + 1} ∪ {0}. Je = {1 ≤ n ≤ Ne : σ −1 (ξne ) > σ −1 (ξn−1 e , the curve visits other new circles Then n is a nonzero element of Je if, after γ visits ξn−1 in Ξ before ξne . That is, there is some ξ ∈ ∪e =e Ξe such that τξ e < τξ < τξ e . Order the n n−1 elements of Je by 0 = se (0) < se (1) < · · · < se (Me ), where Me = |Je | − 1 is the number of times that the progress of γ through Ξe is interrupted. Define se (Me + 1) = Ne + 1. Using this framework, each Ξe can be partitioned into Me + 1 subsets Ξ(e,j) = {ξne : se (j) ≤ n ≤ se (j + 1) − 1}, 0 ≤ j ≤ Me . These are the elements of Ξe which are visited without interruption. Let I = {(e, j) : e ∈ E, 0 ≤ j ≤ Me }. Then {Ξi }i∈I is a finer partition with the desired properties. For i ∈ I, let i = (ei , j). We need to do some relableing. Let e zi = zei , yi = 1 − |zi |, min{Ξi } = ξs i (j) , ei Pyi (Rmax{Ξ } ) e i , max{Ξi } = ξs i (j+1)−1 , Pi = ei Pyi (Rmin{Ξ } ) i 45 where Rmax{Ξ } is the radius of max{Ξi }, and Rmin{Ξ } is the radius of min{Ξi }. By Lemma i i 9, we can see that P[τmax{Ξ } < ∞|Fτmin{Ξ } ] ≤ CPi . i i (2.14) For e ∈ E, let Pe = Pye (re )/Pye (Re ). We claim that for each e ∈ E, we have Me P(e,j) ≤ 4αMe Pe . (2.15) j=0 This follows from Lemma 1, and the details are provided in Section 2.7. Observe that |I| = e∈E (Me + 1) is the number of uninterrupted sequences of circles visited by γ under σ. Then σ induces a map σ ˆ : {1, . . . , |I|} → I so that if n1 < n2 , we have max{σ −1 (Ξσˆ (n ) )} < min{σ −1 (Ξσˆ (n ) )}, and n1 = n2 − 1 implies max{σ −1 (Ξσˆ (n ) )} = 1 2 1 min{σ −1 (Ξσˆ (n ) )} − 1. In other words, σ ˆ is the order that the families Ξi are visited. In 2 particular, we can rewrite the event E σ as E σ = {τmin Ξ σ ˆ (1) < τmax Ξ σ ˆ (1) < · · · < τmin Ξ σ ˆ (|I|) < τmax Ξ σ ˆ (|I|) < ∞}. We will use Theorem 4 to estimate the probability of this event. Fix some e0 ∈ E, and let nj = σ ˆ −1 ((e0 , j)) for 0 ≤ j ≤ Me0 . In the event E σ , the family Ξ(e ,j) is the nj -th family to be visited by the curve. For 0 ≤ j ≤ Me − 1, nj+1 ≥ nj + 2 0 since at least one other family not contained in Ξe0 must be hit by the curve between them. Fix 0 ≤ j ≤ Me0 − 1, and let m = nj+1 − nj − 1 ≥ 1. We are going to apply Theorem 4 to the family • ξˆ0 = min Ξe0 , ξ0 = max Ξ(e ,j) = max Ξσˆ (n ) , and ξ0 = min Ξ(e ,j+1) = min Ξσˆ (n ) 0 0 j j+1 46 • ξˆk = min Ξσˆ (n +k) , and ξk = max Ξσˆ (n +k) for 1 ≤ k ≤ m. j j In plain words, m is the number of other families {Ξi } first visited by γ between first visits of the j-th level and the (j + 1)-st level of the family Ξe0 . The curves ξk , ξˆk are the k-th family first visited before returning to Ξ(e ,j+1) . 0 σ by Define an event by E[max Ξσ ˆ (nj ) ,min Ξσ ˆ (nj+1 ) ] . {τξ0 < τξˆ < τξ1 < · · · < τξm < τξ } ∈ Fτmin Ξ 1 σ ˆ (nj+1 ) 0 Theorem 4 implies that σ P[E[max |Fmax Ξ ] Ξσ σ ˆ (nj ) ˆ (nj ) ,min Ξσ ˆ (nj+1 ) ] ≤ α C m 4− 4 (se0 (j+1)−1) nj+1 −1 Pσˆ (n) . n=nj +1 Varying j = 0, 1, . . . , Me0 − 1 and using inequality (2.14), we see that P[E σ ] ≤ −(α/4) C |I| 4 Me 0 j=1 se0 (j)−1 Pi . i∈I If we use inequality (2.15) and |I| = e Me + 1, we can deduce that the right hand side is bounded above by C |E| C −(α/4) e∈E Me 4 Me 0 j=1 se0 (j) Pe . e∈E Recall that this estimate was based on a fixed e0 ∈ E, but the left hand side does not depend on this choice. Taking the geometric average with respect to e0 ∈ E, we can see that P[E σ ] ≤ C |E| C − α e∈E Me 4 4|E| e∈E Me s (j) j=1 e Pe . e∈E 47 Note that the finer partition I and associated terms Me , se (j) are dependent on our initial choice of order σ. Using the fact that E = ∪E σ and the above inequality, we get  P[E] ≤      C |E| |S(Me ,(se (j))) |C e Me ;(se (j))M j=1 − α e∈E Me 4 4|E| e∈E Me s (j)  j=1 e    Pe , e∈E e∈E (2.16) where S(Me ,se (j))) = {σ ∈ S : Meσ = Me , sσe (j) = se (j), for 0 ≤ j ≤ Me , e ∈ E} e and the first sum in inequality (2.16) is over all possible Me ; (se (j))M j=1 e∈E . That is, for each e, all possible Me ≥ 0 and possible orderings 0 = se (0) < se (1) < · · · < se (Me ) ≤ Ne . Recall that Ne = |Ξe | − 1 is fixed with the initial partition E. To finish the proof of the theorem, it suffices to show that the large term in the parenthesis in (2.16) can be bounded above by some finite constant depending only on |E| and κ. We claim that |S(Me , se (j))| ≤ |E| e∈E Me +1 . (2.17) Notice that the pair (Me , se (j)) completely determines the partition {Ξi }i∈I σ , and the partition σ can be recovered from the induced partition σ ˆ : {1, . . . , |I σ |} → I σ . This is because the order that circles in any given Ξi are visited is predetermined by the ordering on Ξ, and so moving from σ ˆ to σ requires no new information. Next, we claim that σ ˆ is determined by knowing eσˆ (n) , for each 1 ≤ n ≤ |I σ |. If eσˆ (n) = e0 , then σ ˆ (n) = (e0 , j0 ) where j0 can be determined by j0 = min{0 ≤ j ≤ Me0 : (e0 , j) : (e0 , j) ∈ /σ ˆ (m), m < n}. There are |I σ | = ˆ (n) , e∈E (Me +1) terms to determine for eσ 48 each of which has at most |E| possibilities, which proves (2.17). Using this to estimate the constant in (2.16), we see that α − |S(Me ,(se (j))) |C e∈E Me 4 4|E| e Me ;(se (j))M j=1 e∈E α − |E| e∈E (Me +1) C e∈E Me 4 4|E| ≤ e Me ;(se (j))M j=1 Me s (j) j=1 e − α (C|E|)Me 4 4|E| e Me ;(se (j))M j=1 Ne = |E||E| Me s (j) j=1 e e∈E e∈E |E||E| = Me s (j) j=1 e e∈E e∈E e∈E Me s (j) j=1 e . α − 4 4|E| (C|E|)Me e∈E Me =1 0=se (0)<··· 1. 49 2.6 Main theorems The strategy for proving Theorem 1 is to construct a family of circles Ξ and a partition {Ξe }e∈E satisfying the hypothesis of Theorem 5 from the discs {|z − zj | ≤ rj } and the distances lj . The constant given will depend on the size of the partition E, but then it can be shown that that |E| can be bounded above in a way which depends only on the number of points n. Proof of Theorem 1. We can assume without loss of generality that for each j = 1, . . . , n, h the radius rj satisfies rj = lj /4 j for some integer hj ≥ 1. This is because the ratio must −hj −1 satisfy 4 −hj ≤ rj /lj ≤ 4 for some integer hj , and any increase of rj by at most a factor of 4 only affects the constant for each term j. Moreover, if hj ≤ 0, then the j-th term in the inequality (2.1) is equal to 1, so we can assume that that hj ≥ 1. Construct Ξ: for each 1 ≤ j ≤ n, construct a sequence of circles by lj ξjs = {|z − zj | = s }, for 1 ≤ s ≤ hj . 4 The family of circles {ξjs }j,s may not be disjoint, so we may have to remove some circles. For any fixed k ≤ n, let Dk = {|z − zk | ≤ lk /4}, which is a closed disc containing all of the circles centered at zk . For j < k ≤ n, define Ij,k = {ξjs : 1 ≤ s ≤ hj , and ξjs ∩ Dk = ∅}. These are the circles centered at zj which intersect Dk if zk is closer to zj than zj is to {0, 1, z1 , . . . , zj−1 }. Define the family Ξ = {ξjs : 1 ≤ j ≤ hj , 1 ≤ s ≤ hj }\ ∪1≤j j, the number of vertices removed is at most 1. Consequence b) says that for each k > j, the number of edges removed is at most 2. Thus, for each 1 ≤ j ≤ n, the number of components created is bounded above 52 by k>j (1 + 2) = 3(n − j). It follows that |Ej | ≤ 1 + 3(n − j). Summing over j yields n |E| = n |Ej | ≤ j=1 1 + 3(n − j) = n + j=1 3n(n − 1) . 2 This completes the proof that C|E| can be bounded above by some constant Cn < ∞ depending only on n and κ. Final estimate: To finish the proof, we need to show that Pyj (re ) e∈Ej Pyj (Re ) ≤C Pyj (rj ) Pyj (Rj ) . We introduce some new notation. For any annulus A = {z : r ≤ |z − z0 | ≤ R}, where y0 = 1 − |z0 | for z0 ∈ D fixed, let P (A) = Py0 (r)/Py0 (R). For 1 ≤ s ≤ hj , let Aj,s = {lj /4s ≤ |z − zj | ≤ lj /4s−1 }, and let Sj = {s ∈ {1, . . . , hj } : Aj,s ⊂ ∪e∈Ξj Ae }. Using this new notation, what we want to show is that Pyj (re ) e∈Ej Pyj (Re ) P (Aj,s ) ≤ Cn = s∈Sj Pyj (rj ) Pyj (Rj ) hj = Cn P (Aj,s ). s=1 By the estimate in Lemma 1, for each e ∈ Ej , we have Pyj (re )/Pyj (Re ) ≥ (re /Re )α = 4−α . Let Sjc = {1, . . . , hj }\Sj . Then P (Aj,s ) = s∈Sj hj s=1 P (Aj,s ) s∈Sjc P (Aj,s ) ≤ Pyj (rj ) α|S c | 4 j. Pyj (Rj ) We need to estimate |Sjc |. If s ∈ Sjc , then either s = 1 or there is some k > j so that Dk ∩ Aj,s = ∅. By consequence b) of (2.20), for each k > j, this happens at most twice for 53 each k > j. Therefore, |Sjc | ≤ 1 + k>j 2 = 1 + 2(n − j). By equation (2.19), n P[∩ξ∈Ξ {γ ∩ ξ = ∅}] ≤ C(n+(3/2)n(n−1)) j=1 Pyj (rj ) Pyj (Rj ) 4α(1+2(n−j)) = Cn 2 4α(n ) n j=1 Pyj (rj ) Pyj (Rj ) . Using Theorem 1, the analogous estimate for whole-plane SLE can be proven. The strategy will be to appeal to reversibility, reduce to the radial case, and then carefully use the Growth Theorem to show that the estimates from the radial case are comparable to the desired upper bound. Proof of Theorem 2. By the reversibility of whole-plane SLEκ for κ ≤ 8 ([40],[21]), the desired probability is the same as the corresponding probability when γ ∗ is a whole-plane SLEκ from ∞ to 0, which we will assume for the rest of the proof. For any stopping time T , the path t → γ ∗ (T + t), conditioned on γ ∗ (−∞, T ), is a radial SLEκ path in C\γ ∗ (−∞, T ] from γ ∗ (T ) to 0. For z1 , . . . , zn ∈ C\{0}, let z0 = 0, and define R = max{|zj −zk | : 0 ≤ j, k ≤ n}. Assume without loss of generality that rk ≤ lk for each k. Otherwise, the corresponding factor on the right hand side of the estimate is 1. Let T = inf{t > −∞ : |γ ∗ (t)| = 4R}, which is finite a.s. since the radial SLEκ tip converges to the target point with probability 1. Let φ be a conformal map from C\γ ∗ (−∞, T ] to D which fixes 0 and sends γ ∗ (T ) to 1. Then, conditioned on γ ∗ (−∞, T ], the path γ := φ(γ ∗ (T + ·)) is a radial SLEκ trace from 0 to 1. Therefore, ∗ ∗ ∗ P[∩n k=1 {dist(zk , γ ) < rk }|γ (−∞, T ]] = P[∩k {γ ∩ φ(B(zk , rk )) = ∅}|γ (−∞, T ]]. (2.21) 54 By the Schwarz lemma, for each k, zkT := φ(zk ) satisfies |zkT | ≤ 1/4, since |zk | ≤ R and φ takes 4RD into D. Let rkT = max{|φ(z) − φ(zk )| : z ∈ ∂B(zk , rk )}, and define lkT = min{|zk − 1|, |zkT − zjT | : 0 ≤ j < k}. Since |zkT | ≤ 1/4, lkT is not |zkT − 1| for any k, and lkT ≤ 1/2 < 3/4 ≤ 1 − |zkT | := ykT . Therefore, P T (rkT ) y k = P T (lkT ) y k rkT lkT for each k = 1, . . . , n. Theorem 1 and (2.6) imply that (2.21) is bounded above by n T P[∩n k=1 {dist(zk , γ) ≤ rkT }|γ ∗ (−∞, T ]] ≤C k=1 rkT lkT 2−d . (2.22) It suffices to show that the quotients rkT /lkT are uniformly comparable to rk /lk for each k. For each k, we will apply the Growth theorem on the ball B(zk , 3R), since φ is conformal on B(0, 4R). There is a small subtlety, in that φ may not preserve the order of z1 , . . . , zn . That is, for a given k, if k ∗ ∈ {0, . . . , k − 1} so that lk = |zk − zk∗ |, it may be that lkT = |zkT − zjT | for some j = k ∗ . This will only affect our estimate by a constant, since if we define ˜lk = |zkT − zkT∗ |, we have by (2.6) ˜lk |zk − zk∗ | ≤ · |zk − zj | lkT |z −z | 1 + k3R j 2 (1 + (1/3))2 ≤ 1 · = C. 2 (1 − 1/3)2 |zk −zk ∗ | 1 − 3R The second inequality follows by the choice of k ∗ , and since |zk − zj | ≤ R for all 0 ≤ j, k ≤ n. Therefore, for each k, we have rkT /lkT ≤ C(rkT /l˜k ). By applying (2.6) to zk and 55 w ∈ ∂B(zk , rk ) with rkT = |φ(zk ) − φ(w)|, and to lk = |zk − zk∗ |, we get lk 1 + 3R rkT 2 r rk (1 + (1/3))2 r ≤ k· ≤ = C k. 2 2 ˜lk rk lk lk (1 − 1/3) lk 1 − 3R Therefore, we have rkT /lkT ≤ Crk /lk uniformly. Proof of Theorem 3. First, we prove the theorem for radial SLE. Recall that we defined Contd (E; r) = rd−2 Area{z ∈ C : dist(z, E) < r} for any r > 0. Fixing r > 0, we have E[Contd (γ; r)n ] = E[rn(d−2) (Area{z ∈ D : dist(z, γ) < r})n ] n = rn(d−2) E I{z:dist(z,γ) lj , then one of the rd−2 ’s from the outside product is added to the product, and is smaller than ljd−2 . Thus, if f (z1 , . . . , zn ) = n d−2 k=1 min{|zk |, |zk − 1|, |zk − z1 |, . . . , |zk − zk−1 |} and r > 0, then E[Contd (γ; r)n ] ≤ Cn Dn f (z1 , . . . , zn )dA(z1 ) . . . dA(zn ). By Fatou’s Lemma, E[Contd (γ)n ] ≤ lim inf E[Contd (γ; r)n ] ≤ Cn r→0 Dn f (z1 , . . . , zn )dA(z1 ) . . . dA(zn ). If we can show that f is integrable, then we are done. Fix k = 1, . . . , n, and let z1 , . . . zk−1 ∈ D be arbitrary. Let z−1 = 0 and z0 = 1. Then for any k, k−1 min{|zk |, |zk − 1|, |zk − z1 |, . . . , |zk − zk−1 |}d−2 dA(zk ) |zk − zj |d−2 dA(zk ). ≤ j=−1 D D Note that for each j, D − zj ⊂ 2D, and so the above inequality satisfies 2 |z|d−2 dA(z) = (k + 1)2π ≤ (k + 1) 2D rd−1 dr < ∞. 0 By Fubini’s theorem, it follows that Dn f dA . . . dA < ∞, and so we are done in the radial case. The proof for whole-plane SLEκ follows similarly. We start off by writing E[Contd (γ ∗ ∩D); r)n ] = Dn rn(d−2) P[dist(z1 , γ ∗ ) < r, . . . , dist(zn , γ ∗ ) < r]dA(z1 ) . . . dA(zn ). 57 The function f (z1 , . . . , zn ) excludes the distance from 1 in the whole-plane case. That is, n min{|zk |, |zk − z1 |, . . . , |zk − zk−1 |}d−2 . f (z1 , . . . , zn ) = k=1 Then E[Contd (γ ∩ D)n ] ≤ Dn f (z1 , . . . , zn )dA(z1 ) . . . dA(zn ). If D ⊂ RD for R < ∞, then D − zj ⊂ 2RD for each zj ∈ D, so we can perform the same bound as in the radial case, except integrating over 2RD rather than 2D. 2.7 Technical lemmas The following lemma is proven in [23], and serves as a technical lemma to prove that a certain random variable is a stopping time. Lemma 11. Let D ⊂ C be a simply connected domain, and let ρ be a crosscut in D. Let w0 , w1 , and w∞ be connected subsets or prime ends of D such that D\ρ is a neighborhood of all of them in D. Suppose that ρ disconnects w0 from w∞ in D. Let γ(t), 0 ≤ t ≤ T, be a continuous curve in D with γ(0) ∈ ∂D. Suppose for 0 ≤ t < T, D\γ[0, t] is a neighborhood of w0 , w1 , and w∞ in D, and w0 , w1 ⊂ Dt := D(γ[0, t]; w∞ ). For 0 ≤ t < T , let ρt be the first subcrosscut of ρ in Dt that disconnects w0 from w∞ as given by Lemma 2. For 0 ≤ t < T , let f (t) = 1 if w1 ∈ Dt (ρt ; w∞ ), and f (t) = 0 if w1 ∈ Dt∗ (ρt ; w∞ ). Then f is right continuous on [0, T ), and left continuous at those t0 ∈ (0, T ) such that γ(t0 ) is not an end point of ρt0 . We can provide the complete rigorous details for some of the steps in the main theorems, whose arguments would clutter the presentation of the proofs. 58 Proof of (2.10) in Theorem 4: First, I need to show that, in the event E, if t ≤ τj , then D∗ (ξj ) is either contained in Dt (ρt ) or Dt∗ (ρt ) for j = 1, . . . , m. If t ≤ τj , and E occurs, then ξj cannot be swallowed by the curve, since γ still needs to intersect ξj . Therefore, it must be that ξj ⊂ Dt . Moreover, since D(ξj ) is disjoint from the curve ρ, it cannot be that the subcrosscut ρt goes through D(ξj ). It follows that either D(ξj ) ⊂ Dt (ρt ) or D(ξj ) ⊂ Dt∗ (ρt ). Similarly, in the event E, it must be that ξ0 ⊂ Dτm . Let I be totally ordered by (0, 1) < (1, 1) < (1, 2) < (2, 2) < · · · < (m, m) < (m, m + 1). Define a family of events by Ei = E\ ∪i >i Ai . Using a reverse induction argument, we can show that Ei ⊂ {D∗ (ξi2 ) ⊂ Dτ∗ (ρτi )} i1 1 for all i = (i1 , i2 ) = (m, m + 1). Note that E(m,m+1) = ∅. For the base case of the induction, we consider E(m,m) = E\A(m,m+1) = E ∩ {T ≤ τm } ∪ {Dτm (ξm ) ⊂ Dτm (ρτ ) } . m Note that E ∩ {T ≤ τm } = ∅, and the preceding paragraph shows that E ∩ {Dτm (ξm ) ⊂ Dτm (ρτm )} is contained in {Dτm (ξm ) ⊂ Dτ∗m (ρτm )}. This proves the base case. To complete the induction, assume that E(j,j) ⊂ {D∗ (ξj ) ⊂ Dτ∗ (ρτj )}. Then j E(j−1,j) = E(j,j) ∩ Ac(j,j) ⊂ 59 ⊂ {D∗ (ξj ) ⊂ Dτ∗ (ρτj )} ∩ {D∗ (ξj ) ⊂ Dτ∗ (ρτj−1 )} ∪ {D∗ (ξj ) ⊂ Dτj (ρτj )} . j j−1 The intersection with the right hand side is empty, and so we get that E(j−1,j) is contained in {D∗ (ξj ) ⊂ Dτ∗ j−1 (ρτj−1 )}. This completes the proof that the (j, j) case implies the (j − 1, j) case. The argument for (j, j + 1) implies (j, j) is identical, and so the the induction is complete. The final step of the above induction sequence implies that E\ ∪i>(0,1) Ai = E(0,1) ⊂ {D∗ (ξ1 ) ⊂ Dτ∗0 (ρτ0 )} ⊂ A(0,1) , from which claim (2.10) follows. Proof of (2.13) in Theorem 4. Let x = Rj /rj , then N = ln(x) . We are trying to show xα/N 1 − x−α/2 1 − x−α/2N ≤ eα 1 − e−α/4 . If 1 < x ≤ e, then N = 1, and the left hand side is equal to xα 1 − x−α/2 1 − x−α/2 = xα ≤ eα ≤ eα 1 − e−α/4 . Suppose that x > e, in which case N ≤ ln(x) + 1, and so 1 − x−α/2(ln(x)+1) ≤ 1 − x−α/2N . The function 1 − x−α/2(ln(x)+1) =1−e −(α/2) ln(x) ln(x)+1 is increasing, and so is bounded from below for x ∈ (e, ∞) at x = e, which yields the bound 60 1 − e−α/4 . Therefore, we get 1 1 − x−α/2N ≤ 1 1 − e−α/4 . Since ln(x) ≤ N = 0, we get that x1/N ≤ x1/ ln(x) = e, and so xα/N ≤ eα . Combining this with the above inequality completes the proof. Proof of (2.15) in Theorem 5. In the total order on Ξe , max{Ξ( e, j)} < min{Ξ(e,j+1) }, and there are no circles between them by construction. By the assumption that sequential radii in Ξe form a geometric sequence, we know that Rmax{Ξ = 4Rmin{Ξ Therefore, (e,j) } (e,j+1) } expanding the product gives Me P(e,j) = j=0 Pe (re ) Pe (Re ) Me −1 j=0 Me −1 Pye (Rmax{Ξ ) (e,j) } Pye (R{min{Ξ (e,j+1) }} ) ≤ Pe   j=0 Rmax{Ξ (e,j) } Rmin{Ξ (e,j+1) } α  = Pe 4αMe , which proves claim (2.15). Note that the above inequality follows from applying Lemma 1 to each term in the product. Proof of consequences a) and b) in Theorem 1: To prove a), suppose that it is false. Then rad(ξj1 ) 1 2 there are distinct ξj , ξj ∈ Ij,k with ≥ 4. By both being in Ij,k , there is a z ∈ ξj1 ∩Dk rad(ξj2 ) and a w ∈ ξj2 ∩ Dk . Then |z − zj |/|w − zj | ≥ 4. Taking the maximum over all z ∈ Dk then implies that maxz∈D {|z − zj |}/|w − zj | ≥ 4. Taking the minimum of this in w ∈ Dk k contradicts (2.20). The proof of b) follows similarly. Suppose b) is false. Then there exist 2 ≤ r1 < r2 < 61 r3 ≤ hj and w1 , w2 , w3 ∈ Dk such that lj lj lj lj lj lj ≤ |w3 − zj | ≤ r −1 ≤ r ≤ |w2 − zj | ≤ r −1 ≤ r ≤ |w1 − zj | ≤ r −1 . r 43 42 41 43 42 41 From this series of inequalities, we get lj /4r1 |w1 − zj | = 4r3 −1−r1 ≥ 4. ≥ |w3 − zj | lj /4r3 −1 This contradicts (2.20) in the same way as in a), proving b). 62 Chapter 3 Decomposition of backward SLE in the capacity parametrization 3.1 Introduction In some cases, the reverse flow of the Loewner equation is easier to study and can be used to answer questions about the regular, or forward, SLE process. Analysis of the reverse flow was used to show existence of the SLEκ trace for κ = 8 [27]. In [8], a multifractal analysis is used to study moments for the backward SLE flow, which is used to provide a new proof of the Hausdorff dimension of an SLE path. The reversibility of the welding [26] has been used to study ergodic properties of the tip of a forward SLEκ in [39]. BSLE and the conformal welding have been coupled with the Gaussian free field and what is called the Liouville quantum zipper [33], where the welding of the real line onto the backward SLEκ traces is seen as the conformal image of gluing random surfaces together. Recently [4], this connection to the Gaussian free field was used to provide a new proof of the existence of the forward SLE trace. In this chapter, we construct a family of functions Ga,b (x, y) with which we use the a,b Girsanov theorem to construct a new measure Px,y under which φ(x) = y almost surely. That is, we condition the process so that the graph of the welding function passes through 63 the point (x, y). In order to do this, we need the notion of a BSLE process with force points, which are introduced in [26]. A process (ft ) which solves the chordal Loewner equation with driving function λ which satisfies the SDE: dλt = √ κdBt + adt bdt + , ft (q1 ) − λt ft (q2 ) − λt λ0 = x0 is called a BSLEκ (a, b) process started from (x0 ; q1 , q2 ), where a, b ∈ R are weights at the force points q1 , q2 ∈ R. This definition can be extended to include more than two force points, which can also be placed in the interior of H, but this is the amount of generality we will need. This process can be constructed by applying the Girsanov theorem, which will be discussed in a later section, and therefore for κ ≤ 4, the backward SLE process with force points also generates a conformal welding on R+ . For any set U contained in the first quadrant, we then average the measures of paths a,b passing through U by integrating Px,y against IU (x, y)Ga,b (x, y). For κ ∈ (0, 4), and for a = b = −4, we then prove a decomposition theorem relating this particular case to the amount of capacity time that the graph of φ spends in U . In order to establish these results, we review the framework established in [37] to study processes with a random lifetime. For all κ ≤ 4, and for a = b = −κ − 4, we also show that the Green’s function Ga,b (x, y) can be realized as the normalized probability the two points x, y are welded together. This is similar to how the Green’s function for the forward SLEκ process captures the normalized probability that the path passes through a given point. 64 3.2 Processes with a random lifetime For this paper, we will need to review the framework introduced in [37] to study stochastic processes with a random lifetime. We will need to introduce the notation which will be used in this paper, and several propositions will be stated without proof. For complete details, the reader can refer to [37]. Define the space Σ = ∪0 0 : f ∈ C[0, t)}. • Killing: For 0 < τ ≤ ∞, define Kτ : Σ → Σ by Kτ (f ) = f|[0,τ ) , where τf = f min{τ, Tf }. • Continuing: This is an operation which glues two functions together. Define subspaces of Σ by Σ⊕ = {f ∈ Σ : Tf < ∞, f (T − ) := limt→T − f (t) ∈ R} and Σ⊕ = {f ∈ Σ : f f (0) = 0}. Then we define ⊕ : Σ⊕ × Σ⊕ → Σ by f ⊕ g(t) =     f (t), 0 ≤ t < Tf .    f (T − ) + g(t − Tf ), Tf ≤ t < Tf + Tg f ˆ : Σ⊕ × Σ⊕ → Σ × [0, ∞) by f ⊕g ˆ = (f ⊕ g, Tf ). • Time marked continuation: Define ⊕ 65 ˆ records the information of where the first function ends and the second begins. Then ⊕ Probability measures on Σ will be the laws of the random driving function for the backward Loewner process. To talk about measures, we need to create a sigma algebra on Σ. First, for t < ∞, define a filtration by Ft = σ {f ∈ Σ : s < Tf , f (s) ∈ U }, s ≤ t, U ⊂ R is measurable . Then define F = ∨0≤t<∞ Ft . Then the operations above are all measurable on (Σ, F). Given any two measures µ, ν on (Σ, F) let µ ⊗ ν denote the product measure. We then define the following measures: • µ ⊕ ν is defined by the pushforward measure ⊕∗ (µ ⊗ ν) on Σ. ˆ is the pushforward measure ⊕ ˆ ∗ (µ ⊗ ν) defined on Σ × [0, ∞). • µ⊕ν We will also have to work with random measures on different spaces, which are called probability kernels. More precisely, suppose (U, U) and (V, V) are measurable spaces. A kernel from (U, U) to (V, V) is a map ν : (U, V) → [0, ∞) such for each u ∈ U , ν(u, ·) : V → [0, ∞) is a measure and for each E ∈ V, the function ν(·, E) : U → [0, ∞) is U-measurable. If µ is a measure on (U, U), then ν is called a µ-kernel if it is a kernel on the µ-completion of (U, U). We say that ν is a finite µ-kernel if ν(u, V ) < ∞ for µ-a.s. u ∈ U, and we say ν is a σ-finite µ-kernel if V = ∪∞ n=1 Fn such that for each n, for µ-a.s. u ∈ U , we have ν(u, Fn ) < ∞. Combining kernels with measures, we have the following operations for measures: • If µ is a σ-finite measure on (U, U) and ν is a σ-finite µ-kernel from (U, U) to (V, V), 66 then we define µ ⊗ ν on U × V by µ ⊗ ν(E × F ) = ν(u, F )dµ(u). E In this case, µ · ν is the marginal measure on V given by µ · ν(F ) = µ ⊗ ν(U × F ). • If ν is a σ-finite measure on V and µ is a σ-finite ν-kernel from (V, V) to (U, U), we ← − define µ ⊗ ν on U × V by ← − µ ⊗ ν(E × F ) = µ(v, E)dν(v). F • If ν is a σ-finite µ-kernel from Σ to (0, ∞), define a measure Kν (µ) on Σ to be the pushforward measure of µ ⊗ ν under the map K : Σ × (0, ∞) → Σ given by (f, r) → Kr (f ). To study the law of a random process over time, the notion of absolute continuity can be extended to what we call local absolute continuity. For 0 ≤ t < ∞, let Σt = {f ∈ Σ : Tf > t} be the set of functions defined at least until time t. Note that ∩t>0 Σt = C[0, ∞). For each t > 0, let Ft ∩ Σt denote the restriction of Ft onto the subspace Σt . If µ, ν are measures on (Σ, F), then we say that ν is locally absolutely continuous with respect to µ if, for every t > 0, ν|Ft ∩Σt is absolutely continuous with respect to µ|Ft ∩Σt . We will use the notation ν µ to mean absolute continuity and ν µ to mean local absolute continuity. The following propositions are in [37], and will be stated without proof: Proposition 2. Let µ be a measure on (Σ, F) which is σ-finite on F0 . Let (Υ, G) be a measurable space. Let ν : Υ × F → [0, ∞] be such that for every v ∈ Υ, ν(v, ·) is a finite measure on F that is locally absolutely continuous with respect to µ. Moreover, suppose that 67 the local Radon Nikodym derivatives are equal to (Mt (v, ·))t>0 , where Mt : Υ × Σ → [0, ∞) is G × Ft measurable for every t ≥ 0. The ν is a kernel from (Υ, G) to (Σ, F). Moreover, if ξ is a σ-finite measure on (Υ, G) such that µ-a.s., Υ Mt (v, ·)dξ(v) < ∞, then ξ · ν µ, and the local Radon-Nikodym derivatives are Υ Mt (v, ·)dξ(v) for 0 ≤ t < ∞. Proposition 3. Let µ be a probability measure on (Σ, F). Let ξ be a µ-kernel from (Σ, F) to (0, ∞) that satisfies Eµ [|ξ|] < ∞. Then Kξ (µ) µ, and the local Radon-Nikodym derivatives are Eµ [ξ((t, ∞))|Ft ] for 0 ≤ t < ∞. Throughout this chapter, we will use Pκ to denote the measure on Σ which gives the law √ of ( κBt )t>0 , where (Bt )t>0 is a standard one dimensional Brownian motion. Then Pκ is supported on C[0, ∞) ∩ Σ⊕ . We will use Eκ to denote expectation under the measure Pκ , and FtB to be the completion of Ft under Pκ for 0 ≤ t. We state two more propositions from [37] without proof. The first extends the Girsanov theorem to a statement about local absolute continuity, and the second extends the strong Markov property of Brownian motion. Proposition 4. Suppose that (Xt )0≤tt} Mt , 0 ≤ t < ∞. (3.2) If (θ)t is an FtB -adapted increasing process, it induces a kernel from Σ to [0, ∞) defined by dθ· (f ), which is the random measure induced by the monotonic function θt (f ) on t ∈ [0, ∞). Proposition 5. Let (θt )0≤t<∞ be a right continuous increasing FtB -adapted process that satisfies θ0 = θ0+ = 0 and Eκ [θ∞ ] < ∞. Then ˆ κ = Pκ ⊗ dθ. Kdθ (Pκ )⊕P Thus, Kdθ ⊕ Pκ (3.3) Pκ , and θ∞ is the Radon-Nikodym derivative. Define ΣL to be the set of λ ∈ Σ which generate a Loewner trace. By the existence of the SLEκ trace for any κ > 0, we know that Pκ is supported in ΣL . Let ΣC = ∪0 0, define ˜ t , and let Gt (z) = F˜t ◦F −1 (z), which maps H\{γt −λt } Ft (z) = ft (z)−λt and F˜t (z) = f˜t (z)−λ t ˜ t }, where (γt )t≥0 and (˜ onto H\{˜ γt − λ γt )t≥0 are the families of backward Loewner traces ˜ respectively. driven by λ and λ For κ < 4, since γt − λt has the same law under Pκ as a forward SLEκ trace up to time t, it was proven in [27] that γt − λt is the boundary of a H¨older domain. In [6], it was shown that this condition is enough to make the image of the trace conformally removable. That is, any conformal map defined on H\{γt − λt } which can be extended continuously to the boundary can be conformally extended to H. Suppose Φ(t) = (bt , at ), and assume −at ≤ −y < 0 < x ≤ bt with φ(x) = y. Note that since both processes have the same welding process, they also have the same welding function φ. Therefore, ft (x) = ft (−y) and f˜t (x) = f˜t (−y), and so Gt extends continuously ˜ t . Conformal removability then implies that Gt extends to a conformal from γt − λt to γ˜t − λ 70 map from H onto H, and so it must be that Gt (z) = az + b cz + d for some a, b, c, d ∈ R with ad − bc = 1. ˜ t . This Assuming Φ(t) = (bt , at ), we get ft (−at ) = ft (bt ) = λt and f˜t (−at ) = f˜t (bt ) = λ implies that Gt (0) = 0, and so b = 0. Also, ft and f˜t each fix ∞, and so c = 0. Therefore, ˜ t , the scaling rule for half Gt (z) = az for some constant a. Since Gt (γt − λt ) = γ˜t − λ ˜ t ) = 2t, and so a = 1 or plane capacity implies that 2t = hcap(γt − λt ) = |a|2 hcap(˜ γt − λ a = −1. Since ft and f˜t both have positive derivative at ∞, we have a = 1. Therefore, ˜ t for each t > 0. Since the reverse flow is uniquely determined by the driving γt − λt = γ˜t − λ ˜ function, we get that λ = λ. 3.3 Glossary of notation for Ito’s formula calculations We define some terms which will be used for Ito’s formula calculations for this chapter in multiple sections. • Fix x, y > 0, and let λt be a driving function for the backward Loewner equation (1.4). Let (ft )t≥0 be the solution of (1.4) driven by λ. For z ∈ R, let τz be the lifetime of (1.4) started from z. If x, y > 0, let τx,y = min{τx , τ−y }. • For t ≤ τx , define Xt = Xt (x) = ft (x) − λt . • For t ≤ τ−y , define Yt = Yt (y) = λt − ft (−y). Note that Xt and Yt are defined specifically to be nonnegative for all t ≤ τx,y . 71 • These two formulas imply that d(Xt + Yt ) −2 = dt. Xt + Yt Xt Yt • For t < τx,y , let Wt = Wt (x, y) = (3.4) Xt − Yt . Then Wt ∈ (−1, 1) for each t < τx,y . If Xt + Yt τx = τ−y , then Wτx,y is well defined and is either 1 or −1. • For t ≤ τx,y , define ut = ut (x, y) = 0t κ dr. By equation (3.4), this is equal to Xr Yr κ u(t) = − ln 2 X t + Yt x+y . (3.5) The process ut will be used to define a random time change. • For Z ∈ {X, Y, W }, we will use Zˆt = Zu−1 (t) . 3.4 Generalized Green’s functions In this section, we will be doing Ito’s formula calculations using the framework in Section 3.3. For the first lemma, λt = √ κBt is the driving function for BSLEκ . We will then do calculations when λt is the driving function for the BSLEκ (a, b) process. For κ ≤ 4, let (ft ) be the standard backward SLEκ process driven by λt = √ κBt . For x, y > 0 define Xt = Xt (x) = ft (x) − λt and Yt = Yt (−y) = λt − ft (−y) as in Section 3.3. Ito’s formula then implies that √ √ 2 2 dt, and dYt = κdBt − dt. dXt = − κdBt − Xt Yt 72 Note that ft (x) is decreasing in t and ft (−y) is increasing in t. Since λt is unbounded in both directions, then τx = inf{t > 0 : Xt (x) = 0} < ∞ a.s. and τ−y = inf{t > 0 : Yt (y) = 0} < ∞ with probability 1, and Xt , Yt track the flow of x and −y towards 0. The stopping time τx,y = τx ∧ τ−y stops the process as soon as either x or −y are absorbed by the welding. What we want to do is weight the measure Pκ in such a way that τx,y = τx = τ−y almost surely. That is, we want to change the probability measure so that φ(x) = y with probability a,b 1. We will find a new family of measures Px,y on Σ which do this, and we call the process an extended BSLEκ (a, b) process started from (0; x, −y). a,b Proposition 6. Fix a, b ∈ R. Then Mt a b − − a,b = Mt (x, y) := Xt κ Yt κ (Xt +Yt )γ ft (x)p ft (−y)q is a local martingale if and only if p=− a(a + κ + 4) , 4κ q=− b(b + κ + 4) ab , and γ = − . 4κ 2κ In this case, the process (ft )t>0 weighted by this local martingale using the Girsanov theorem is a backward SLEκ (a, b) process started from (0; x, −y). Proof. For ease of notation, let α = −a/κ and β = −b/κ. This follows from an Ito’s formula calculation, which will be included for completeness. Observe first that ∂t ft (z) = 2ft (z) d −2 ∂t ft (z) = · (−f (z)) = . t dz (ft (z) − λt )2 (ft (z) − λt )2 Therefore, dft (z)r df (z) 2r =r t = dt. r ft (z) ft (z) (ft (z) − λt )2 73 Applying this at both x and −y gives dft (x)p dft (−y)q 2p 2q = dt and = 2 dt. p q 2 ft (x) ft (−y) Xt Yt Also, direct applications of Ito’s formula gives √ dXtα −α κ h(α) = dBt + dt, α Xt Xt Xt2 d(Xt + Yt )γ −2γ dt, = γ (Xt + Yt ) Xt Yt √ α κ h(β) = dBt + 2 dt, β Yt Yt Yt β dYt where h(t) = −2t + t(t − 1)κ . By the product rule, we get that 2 β β d Xtα , Yt dXtα dYt dMt = + β + β Mt Xtα Yt Xtα Yt β = Notice that d(Xt + Yt )γ dft (x)p dft (−y)q + + + (Xt + Yt )γ ft (x)p ft (−y)q β d Xtα , Yt dXtα dYt −2γ 2p + + + + 2+ = α β β Xt Xt Yt Xt Yt Xtα Yt √ √ −α κ β κ h(α) + 2p h(β) + 2q + + − dBt + Xt Yt Xt2 Yt2 2q Yt2 dt αβκ + 2γ Xt Yt dt. a b c + − = 0 for all x, y, ∈ R if and only if a = b = c = 0. This occurs if x2 y 2 xy and only if p, q, and γ are chosen as indicated. Once these are chosen, we then get that dMt = Mt √ √ −α κ −β κ + Xt Yt dt. Using this as a drift term and weighting the original measure to obtain a new probability 74 ˜ the Girsanov theorem says that measure P, dλt = √ κdBt = √ ˜t + κdB −ακ −βκ + ft (x) − λt ft (−y) − λt dt, ˜ ˜t is a P-Brownian where B motion. This is precisely the definition of a backward SLEκ (−ακ, −βκ) process started from (0; x, −y), where a = −ακ and b = −βκ. The force points at x and −y serve as magnets. If a and b are chosen correctly, we can force x and −y to be absorbed by the process at the same time, which would imply that φ(x) = y. Along with Xt and Yt , define Wt as in Section 3.3 by Wt = X t − Yt , which is X t + Yt a process that stays in (−1, 1) for as long as it exists. It only fails to exist if τx = τ−y , where τz is the time at which z is swallowed by the BSLE traces. If this does not happen (which is a.s. under Pκ ), then Wt is constant at ±1 after τx,y . The strategy is to perform an appropriate random time change, under which W becomes a diffusion process. We can then analyze the lifetime of this diffusion. Lemma 13. A BSLEκ (a, b) process started from (0; x, −y) welds x and y together with probability 1 if a, b ≤ −4 − κ . 2 Proof. Consider a backward chordal SLE(κ; a, b) process started from (0; x, −y) with driving function λ. Let Xt , Yt , Wt , and u(t) be defined as in Section 3.3. We proceed similarly to Proposition 6. By definition, √ dXt = − κdBt − a b − X t Yt dt − √ 2 dt, and dYt = κdBt + Xt 75 a b − Xt Yt dt − 2 dt. Yt Therefore, √ dXt = − κBt − a+2 b − Xt Yt dt, and dYt = √ κdBt + a b+2 − Xt Yt dt. (3.6) Simple Ito’s formula calculations show that dXt + Yt −2 dXt − Yt 2 2 = dt, and = dt − dλt , Xt + Yt X t Yt Xt − Yt Xt Yt Xt − Yt and therefore √ 4 2 −2 κdBt dWt = − dλt =⇒ dWt = + Wt Xt Yt Xt − Yt Xt + Yt Using the basic algebraic equality 2 4Wt bXt − aYt +2 X t Yt (Xt + Yt )Xt Yt dt. x−y bx − ay = (b − a) + (b + a) , this becomes x+y x+y √ −2 κdBt dWt = + Xt + Yt 4Wt (b − a) + (b + a)Wt + Xt Yt Xt Yt dt. κ κ dr = − ln((Xt + Yt )/(x + y)).Using u to perform a Recall the function u(t) = 0t X r Yr 2 random time change, we get ˜t = − dW ˜ ˜ 2 dB ˜t + (b − a) + (b + a)Wt dt, 1−W t κ (3.7) ˜ t = W −1 . Anything satisfying equation (3.7) is where B˜s is a Brownian motion and W u (t) associated with a radial Bessel process, and is studied in [39]. Thus, our process W is such a process traversed at a random speed. Observe that uτx,y is the lifetime of (3.7). Also, note that Xτx,y + Yτx,y = 0 if and only if uτx,y = ∞. According to the appendix in [39], the 76 lifetime of (3.7) is infinite if a, b ≤ −2 − κ κ with probability 1. Hence, if a, b ≤ −2 − , then 2 2 Xτx,y + Yτx,y = 0, and thus x and y are welded together. Proposition 6 and Lemma 13 motivate us to define the a, b-backward SLEκ Green’s ab b a function for a, b ≤ −2 − κ/2 by Ga,b (x, y) = x− κ y − κ (x + y)− 2κ , so that M a,b (x, y) = Ga,b (Xt , Yt )ft (x)p ft (−y)q is the local martingale whose weighting gives the BSLEκ (a, b) process starting from (0; x, −y). What we want to do now, however, is run the backward SLEκ (a, b) process until it welds x and y together, and then revert the driving function back to a standard Brownian motion after τx,y . a,b To make this precise, let a, b ≤ −2 − κ/2. Let Px,y on Σ be the measure of the driving function for the BSLEκ (a, b) process starting from (0; x, −y) until time τx,y . Then τx,y < ∞ a,b means that Px,y is supported on Σ⊕ . Since Pκ , the law of √ κ times a Brownian motion, is a,b supported on Σ⊕ , that means we can define the measure Px,y ⊕ Pκ on ΣW , which we will call the law of the driving function for an extended BSLEκ (a, b) process started from (0; x, −y). a,b Then W∗ (Px,y ⊕Pκ ) is the measure on ΣC giving the law of the extended BSLEκ (a, b) started from (0; x, −y) welding curve. Under this measure, the welding curve passes through (x, y) almost surely. Our goal is to show that backward SLE can be decomposed into the average of weighted extended BSLEκ (a, b) processes. Fix a measurable U ⊂ Q1 = [0, ∞) × [0, ∞), and observe a,b that the measure Px,y is a probability kernel from Q1 to (Σ, F). Define a new measure on (Σ, F) by a,b PU = a,b U Px,y Ga,b (x, y)dxdy. We say that BSLEκ admits a BSLEκ (a, b) decomposition if for all measurable U ⊂ Q1 77 there is some increasing random process Θ = ΘU t t>0 so that a,b ˆ U PU ⊕P κ = Pκ ⊗ dΘ· . Recall then that dΘU · is a kernel from Σ to [0, ∞). This is a backward analogue to the definition in [37], which defines what it means for SLEκ to admit an SLEκ (ρ) decomposition. In [37], it is proven that SLEκ admits both an SLEκ (κ− 8) decomposition and an SLEκ (−8) decomposition. In the first case, this corresponds to weighting the SLEκ curve against the natural parametrization, and in the latter the SLEκ curve is weighted by the capacity parametrization. In the next section, we will show that BSLEκ admits a BSLEκ (−4, −4) decomposition, and see that the corresponding weight is capacity time. a,b We will need to study the distribution of the stopping time τx,y under Px,y . a,b Lemma 14. Under Px,y , the stopping time τx,y satisfies τx,y = (x + y)2 ∞ − 4 t ˆ 2 )dt e κ (1 − W t 4κ 0 (3.8) a,b ˆ satisfies (3.7) for a, b ≤ −2 − κ/2. Px,y a.s., where W Proof. Recall the time change from before, where u(t) := 0t κ κ ds = − ln((Xt +Yt )/(x+ Xs Ys 2 y)), and we will use the convention that Zˆt = Zu−1 (t) , for Z ∈ {X, Y, W }. We claim that a,b Px,y -almost surely, ∞ τx,y = 0 ˆ s Yˆs X ds. κ (3.9) 2 a,b Recall that under Px,y ,we have τx = τy , and so u(τx,y ) = − ln(Xτ + Yτ ) = ∞ Px,y -almost κ 78 surely, where τ = τx,y . We are using the fact that x and y are welded together here. Then ∞ 0 u−1 (τ ) X τ τ X Y ˆ s Yˆs ˆ s Yˆs (u−1 ) (s) X t t ds = u (t)dt = 1dt = τx,y . ds = κ κ (u−1 ) (u(u−1 (s))) κ u−1 (0) 0 0 Using W = X −Y 4XY ˆ t + Yˆt = and the identity 1 − W 2 = , and the fact that X X +Y (X + Y )2 2 (x + y)e− κ t , we can rewrite the integrand in equation (3.9) and obtain (3.8). a,b Define a function σ a,b (x, y; t) = Px,y [τx,y ≤ t]. By equation (3.8), we can conclude that σ a,b (x, y; t) = σ a,b x y √ , √ ; 1 , for all t > 0. t t (3.10) This holds because (x + y)2 ∞ − 4 s ˆ 2 )ds ≤ t e κ (1 − W s 4κ 0 if and only if √ √ (x/ t + y/ t)2 ∞ − 4 s ˆ 2 )ds ≤ 1, e κ (1 − W s 4κ 0 √ √ ˆ t )t>0 has the same starting point at (x, y) as at (x/ t, y/ t). where the diffusion (W It is worth observing that the local martingale obtained in Proposition 6 is a local martingale, but not a martingale. Lemma 15. If a, b < 0, then a,b Eκ [Mt (x, y)] = Ga,b (x, y)Px,y (τ(x,y) > t) = xα y β (x + y)γ (1 − σ a,b (x, y; t)), (3.11) where α = −a/κ, β = −b/κ, and γ = −ab/(2κ). In particular, Mt is a local martingale, but 79 a strict supermartingale. Proof. For this lemma, since x and y are fixed, we will let τ = τx,y . Define τn = inf{t : Mt ≥ n}. Then for each n, Mt∧τn is a bounded local martingale, and hence a martingale. Therefore, Eκ [Mt ] = lim Eκ [Mt I{τn >t} ] = lim Eκ [Mt∧τn Iτn >t ] n→∞ n→∞ a,b a,b = G(x, y) lim Px,y (τn > t) = G(x, y)Px,y (∪n {τn > t}) . n→∞ a,b a,b We claim that Px,y (∪n {τn > t}) = Px,y (τ > t). It is clear that ∪n {τn > t} ⊂ {τ > t}, since τ is the lifetime of (Mt ). It suffices to show ∩n {τn ≤ t} ⊂ {τ ≤ t}. In the event ∩n {τn ≤ t}, Ms is unbounded for s ∈ [0, t]. It can easily be seen that if t ≤ τ , then ft (x) − λt ≤ x − λt ≤ x + y, and λt − ft (−y) ≤ x + y similarly. It follows immediately that β Xtα Yt ≤ (x + y)α+β , and so this factor of Mt is bounded uniformly for fixed x, y. Therefore, if Ms is unbounded on [0, t], then it must be that either (Xs + Ys )γ or fs (x)p fs (−y)q is unbounded on [0, t]. In the first case, since γ = −ab/(2κ) < 0, this implies that Xs + Ys → 0 before time t. That is, τ(x,−y) ≤ t. In the latter case, the reverse Loewner equation implies fs (x)p fs (−y)q = exp{ s 2q 2p + 2 dr}. 2 Yr 0 Xr This term is unbounded only if p and q are nonnegative and Xs , Ys → 0 before time t. Therefore, in either case, we have τ ≤ t. 80 3.5 Capacity parametrization The goal of this section is to show that BSLEκ has a BSLEκ (−4, −4) decomposition, and to show the relationship between this decomposition and the capacity time for the welding −4,−4 curve. For this section, we will use Px,y := Px,y for x, y > 0. Lemma 16. 0∞ 0∞ Px,y [τx,y ≤ 1]dxdy ∈ (0, ∞). This lemma states that the probability of two points which are far away being welded together in a short time decays rapidly. By the scaling relation, what matters most is how relatively far x is from 0 in comparison to y. This is reflected in the proof, where we perform a change of variables from (x, y) to r(t, 1 − t) and perform the most work integrating the single variable t. Proof. By equation (3.8), we have Px,y -almost surely that τx,y ≤ (x + y)2 ∞ − 4 t (x + y)2 e κ dt = . 4κ 16 0 Thus, if x + y ≤ 4, then Px,y [τx,y ≤ 1] ≥ Px,y [τx,y ≤ (x + y)2 /16] = 1, and therefore ∞ 0 ∞ 0 Px,y [τx,y ≤ 1] ≥ 1dxdy > 0. x+y≤4 To use the scaling relation for BSLE to show that 0∞ 0∞ Px,y [τx,y ≤ 1]dxdy < ∞, we perform a change of variables. Observe that the distribution of τx,y is the same as x . To simplify notation, we will write Pt = Pt,1−t and (x + y)2 τt,1−t , where t = x+y 81 τt = τt,1−t for 0 < t < 1. Performing the change of coordinates x = rt, y = r(1 − t), we get that ∞ 0 ∞ 0 ∞ 1 Px,y [τx,y ≤ 1]dxdy = 0 0 1 Pt [τt ≤ 2 ]rdrdt. r For any fixed t, observe that ∞ 0 Pt [τt ≤ r−2 ]rdr = ∞ 0 Pt 1 1 1 ≥ r2 rdr = Et , τt 2 τt where Et is the expectation with respect to the measure Pt . Thus, it suffices to show that 1 1 0 Et τ dt is finite. t For any small δ > 0, we can break up the integral as 1 0 Et δ/2 1−δ/2 1 1 1 1 1 dt = Et dt + Et dt + Et dt. t τt τt τt 0 δ/2 1−δ/2 To complete the proof, we will produce a small δ > 0 so that: a) There is a constant Cδ < ∞ so that if 0 < t ≤ δ/2, Et 1 1 ≤ Cδ Eδ/2 τt τδ/2 < ∞. (3.12) b) For each t ∈ (δ/2, 1 − δ/2), Et 1 1 1 ≤ Eδ/2 + E1−δ/2 . τt τδ/2 τ1−δ/2 (3.13) Note that if (3.12) is proven, a symmetric argument proves a similar bound for t ∈ [1−δ/2, 1). Once a correct δ can be chosen for (3.12), then (3.13) and both boundary estimates imply 82 that 01 Et [1/τt ] dt < ∞. Note that t0 = x/(x + y) being near 0 is the same as starting the process Wt = (Xt − Yt )/(Xt + Yt ) at w = (x − y)/(x + y) = 2t0 − 1 near −1, and t0 being near 1 corresponds to starting the process Wt near 1. Under the measure Pt0 , equation (3.7) reduces to ˆ 2 dBt + −8 W ˆt = − 1 − W ˆ t dt, dW t κ ˆ 0 = w. W (3.14) We will use this diffusion to prove (3.13). Suppose δ > 0 is small, and fix any t∗ ∈ (δ/2, 1 − δ/2). Let (Wt∗ )t>0 be (3.14) started at w = 2t∗ −1, (Wt1 )t>0 be (3.14) started at w = −1+δ, and let (Wt2 )t>0 be (3.14) started at w = 1 − δ. Then for each t > 0, j 1 − (Wt∗ )2 ≥ min{1 − (Wt )2 : j = 1, 2}. Thus, by equation (3.8), we can see that there is a coupling so that τt∗ ≥ min{τδ/2 , τ1−δ/2 }, where τt∗ , τδ/2 , and τ1−δ/2 are defined on the same probability space, and each τw has the distribution of τw under Pw for w ∈ {δ, 1 − δ/2}. Therefore, 1 1 1 1 1 ≤ max{ , }≤ + . τt∗ τδ/2 τ1−δ/2 τδ/2 τ1−δ/2 Applying the expectation gives (3.13). The rest of the proof is showing that (3.12) holds. Fix (for now) some δ > 0, and let −1 < w < −1+δ. In the t coordinates, suppose W (0) = w = 2t−1, so that −1 < w < −1+δ corresponds to 0 < t < δ/2. Fix any t∗ ∈ (0, δ/2), and let w = 2t∗ − 1 be the starting point ˆ t )t>0 solving (3.14), where W ˆ has an infinite lifetime. Define a stopping of the process (W 83 ˆ t = −1 + δ}. Then equation (3.8) implies time by Tδ = inf{t > 0 : W τt∗ = ∞ ∞ 4 4 1 ˆ 2 )dt ≥ 1 ˆ 2 )dt e− κ t (1 − W e− κ t (1 − W t t 4κ 0 4κ T δ 4 = e− κ Tδ ∞ 4 1 ˆ 2 )dt. e− κ t (1 − W t+Tδ 4κ 0 The right hand side is the stopping time for the process started at t = δ/2, so conditioned 4 on {Tδ = s}, we have τt∗ ≥ e− κ s τδ/2 in distribution. Therefore, Et∗ [1/τt∗ |Tδ = s] ≤ 4 e κ s Eδ/2 1/τδ/2 . It follows that Et∗ 1 = τt∗ ∞ 1 |T ∈ [n − 1, n) Pt∗ [n − 1 ≤ Tδ < n] τt∗ δ Et∗ n=1 ∞ 1 4 e κ n+1 Eδ/2 ≤ n=0 τδ/2 Pt∗ [Tδ ≥ n]. We claim that Pt∗ [Tδ ≥ n] decays exponentially in n. Observe that ˆ t < −1 + δ}, . . . , { Pt∗ [Tδ ≥ n] = Pt∗ [{ sup W 0≤t≤1 sup n−1≤t≤n ˆ t < −1 + δ}]. W ˆ is found, and it can be extended via the Kolmogorov In [39], the transition density for W consistency theorem to a process which can start at −1. Observe then that during each time interval k − 1 ≤ t < k, the process lies above a coupling of the same diffusion started ˆ started at −1, then at −1. Therefore, if pδ = P0 [Tδ ≥ 1], where P0 is the law of W ˆ t ≤ δ − 1] ≤ pδ . Moreover, by the Markov property, these couplings are Pt∗ [supk−1≤t 0. κ δ δ Let T = inf{t : θt = }, and T = inf{t : θt = π − }, so that T = min{T , T }. Then 2 2 where A = Pδ/2 [T ≤ ] = Pδ/2 [T ≤ , T = T ] + Pδ/2 [T ≤ , T = T ]. Both of these probabilities will be estimated similarly, so we will show that the first has sufficiently strong decay. Observe that Pδ/2 [T ≤ , T = T ] ≤ Pδ/2 [ inf θt ≤ δ /2, sup θt < π − δ /2] 0≤t≤ 0≤t≤ = Pδ/2 [ inf θt − δ ≤ −δ /2, sup θt ≤ π − δ /2] 0≤t≤ 0≤t≤ ≤ P[ inf Bt − At ≤ −δ /2] = P[ sup Bt + At ≥ δ /2] ≤ P[ sup Bt ≥ δ /2 − A ], 0≤t≤ 0≤t≤ 0≤t≤ where P is the law of standard Brownian motion. Note that with δ fixed, can be chosen small enough so that this probability is nontrivial. Moreover, by the scaling property of 86 Brownian motion and the reflection principle, we get =P sup Bt ≥ δ /2 − A √ 0≤t≤1 Using the elementary bound P[N (0, 1) ≥ M ] ≤ Pδ/2 [T ≤ , T = T ] ≤ C √ δ /2 − A √ = 2P B1 ≥ . C −M 2 /2 e , this implies that M exp{− √ ((δ /2) − A )2 } ≤ C e−K/ , 2 where C, K are both constants depending only on δ and κ. An identical bound can be found for the event {T ≤ },so the same bound holds for Pδ/2 [T ≤ ]. Combining the inequalities of the previous paragraphs, we can conclude that 1 Eδ/2 τδ/2 ∞ ≤C n=0 n 1 4 −n−1 1 − e− κ 2 n 2− 2 e−K2 . n The terms e−K2 4 −n−1 1 − e− κ 2 are bounded, and therefore 1 ≤C Eδ/2 τδ ∞ n 2− 2 < ∞. n=0 We will now make use of the results in Section 3.2, which originate from [37]. For ∗ ∗ any N > 0 and x, y > 0, define the measure Px,y;N by Px,y;N [E] = P∗x,y [E\ΣN ], where ΣN = {f ∈ Σ : Tf > N }. Note that this is not a probability measure. By Proposition 4, we 87 know that the local Radon Nikodym derivative P∗x,y with respect to Pκ is given by dP∗x,y |Ft ∩Σt dPκ |Ft ∩Σt = I{τx,y >t} Mt (x, y) = I{τx,y >t} G(Xt (x), Yt (y))ft (x)ft (−y), 4 4 (3.16) 8 where G(x, y) = G−4,−4 (x, y) = x κ y κ (x + y)− κ is the (−4, −4)-backward SLE Green’s function in the capacity parametrization. Let E ∈ Ft ∩ Σt . If t ≥ N , then P∗x,y;N [E] = 0. If t < N , then P∗x,y;N [E] := P∗x,y [E] − P∗x,y [E ∩ ΣN ] = = MN (x, y) Mt (x, y) dPκ − dPκ E∩ΣN G(x, y) E G(x, y) 1 Mt (x, y) − Eκ [MN (x, y)|Ft ]dPκ . G(x, y) E Therefore, we conclude that dP∗x,y;N |Ft ∩Σt dPκ |Ft ∩Σt = I{t 0, and limt→∞ Gt (x, y) = G(x, y)P∗x,y [τx,y < ∞] = G(x, y). Let Cκ,t := 0∞ 0∞ Gt (x, y)dxdy = 0∞ 0∞ G(x, y)P∗ (τx,y ≤ t)dxdy, and let Cκ = Cκ,1 . We claim that Cκ,t = tCκ , where Cκ ∈ (0, ∞). The equality follows from scale invarience 88 and the change of variables formula. In particular, ∞ Cκ,t = 0 ∞ ∞ Gt (x, y)dxdy = 0 ∞ √ = ( t)2 0 0 ∞ 0 G1 x y √ ,√ t t dxdy ∞ 0 G1 (x, y)dxdy. Since Cκ = 0∞ 0∞ G(x, y)P∗x,y [τx,y < 1]dxdy, Lemma 16 and the fact that G(x, y) > 0 implies that Cκ > 0. Moreover, G is bounded, which combined with Lemma 16 implies that Cκ < ∞. Recall that Q1 = [0, ∞) × [0, ∞) denotes the first quadrant. Given a λ ∈ ΣW , we denote the welding curve generated by λ by Φ : [0, ∞) → Q1 , where Φ(t) = (x, y) if x and −y are absorbed at time t. That is, φ(x) = y and τx = τ−y = t. For U ⊂ Q1 measurable, we define an increasing process (ΘU t ) by −1 ΘU t = Cκ m+ {Φ (U ) ∩ [0, t]}, (3.18) where m+ {·} is one dimensional Lebesgue measure on [0, ∞). Then ΘU t is the capacity time spent by Φ in the set U before t. Note that then dΘU · is a kernel from (Σ, F) to [0, ∞) and can be thought of as an F-measurable random measure on the line. Also, recall that (x, y) → Px,y is a probability kernel from Q1 to Σ, and we can define a measure on Σ for −4,−4 U ⊂ Q1 measurable by PU = PU = U Px,y G(x, y)dxdy. Theorem 6. Fix κ ∈ (0, 4). For any measurable U ⊂ Q1 , −4,−4 ˆ PU ⊕Pκ = Pκ ⊗ dΘU · , 89 (3.19) −4,−4 W∗ (Px,y −4,−4 where MU = MU ← − −4,−4 ⊕ Pκ ) ⊗ IU G−4,−4 (x, y)dxdy = W∗ (Pκ ) ⊗ MU , (3.20) is the kernel from the set of welding curves W(ΣW ) to Q1 defined by U MU (Φ, E) = Φ∗ (dΘU · ){E} = dΘ{Φ−1 (E)} . To ensure clarity, we recall that the measures in equation (3.19) are measures on the space Σ⊕ × Σ⊕ , and the measures in equation (3.20) are measures on ΣC × Q1 supported on W(ΣW ) × Q1 . Throughout the rest of the chapter, for any measure µ on Σ, we will use µ|t to denote the restricted measure µ|Ft ∩Σt . The proof of the theorem follows the same strategy as the proofs of Theorems 4.1, 5.1, and 6.1 in [37]. Proof. For any t ≤ N , the renewal property of the backward Loewner equation and the Markov property of √ κBt imply that Mt (x, y)−Eκ [MN (x, y)|Ft ] = I{τx,y >t} |ft (x)||ft (−y)|GN −t (Xt , Yt ), where Xt (x) = ft (x) − λt and Yt (y) = λt − ft (−y). Also, recall the welding curve satisfies Φ(t) = (bt , at ), where bt = sup{x > 0 : τx ≤ t} and at = sup{y > 0 : τ−y ≤ t}. Therefore, ∞ 0 ∞ ∞ 0 Mt (x, y)−Eκ [MN (x, y)|Ft ]dxdy = ∞ bt GN −t (Xt (x), Yt (y))|ft (x)||ft (−y)|dxdy ∞ = 0 at ∞ 0 GN −t (x, y)dxdy = Cκ,N −t = (N − t)Cκ . (3.21) The second equality follows from a direct application of the change of variables formula, since ft (−at ) = ft (bt ) = 0. Let ξ be the measure on Q1 given by dξ = G(x, y)dxdy, and observe that ν((x, y), E) = Px,y;N (E) is a kernel from Q1 to Σ. By Proposition 2, ξ · ν = Q Px,y;N G(x, y)dxdy 1 is absolutely continuous with respect to Pκ , and the local Radon-Nikodym derivatives are 90 given by ∞ ∞ dP dξ · ν|t x,y;N |t = G(x, y)dxdy dPκ |t dPκ |t 0 0 ∞ ∞ = 0 0 Mt (x, y) − Eκ [MN (x, y)|Ft ] G(x, y)dxdy = Cκ (N − t) ∨ 0. G(x, y) The second equality comes from (3.17), and the last inequality comes from (3.21). Applying Proposition 3, we get ∞ 0 ∞ Px,y;N G(x, y)dxdy = KCκ d(t∧N ) (Pκ ). 0 Taking N → ∞, it follows that ∞ PQ1 = 0 ∞ 0 Px,y G(x, y)dxdy = KCκ dt (Pκ ). ˆ κ = PQ ⊗P ˆ κ , and so Proposition 5 implies that Therefore, KCκ dt (Pκ )⊗P 1 ˆ κ = Pκ ⊗ Cκ m+ . PQ1 ⊗P (3.22) Note that this is (3.19) in the case U = Q1 . We claim that for any (x, y) ∈ Q1 , ˆ ∗ (Px,y ⊕P ˆ κ ) = W∗ (Px,y ⊕ Pκ ) ⊗ δ(x,y) , W (3.23) where δ(x,y) is the point mass measure on Q1 . Observe that Px,y is supported on {f ∈ ˆ ∗ (Px,y ⊕P ˆ κ ) is the pushforward of the measure ΣW : Wf (Tf ) = (x, y)}, and recall that W ˆ ◦⊕ ˆ ⊕(f, ˆ : Σ⊕ × Σ⊕ → ΣC × Q1 given by W( ˆ g)) = Px,y ⊗ Pκ under the transformation W 91 ˆ ⊕ g, Tf ) = (Φ, Φ(Tf )), where Φ is the welding induced by f ⊕ g. Thus, if E ⊂ ΣC and W(f U ⊂ Q1 , ˆ ∗ (Px,y ⊕P ˆ κ ){E × U } = Px,y ⊗ Pκ {(f, g) : Φ = W(f ⊕ g) ∈ E and Φ(Tf ) = (x, y) ∈ U }, W which proves the claim. Integrating (3.23) over all (x, y) ∈ Q1 with respect to the measure G(x, y)dxdy gives − ˆ ∗ (Pκ ⊗ Cκ m+ ) = W ˆ ∗ (PQ ⊕P ˆ κ ) = W∗ (Px,y ⊕ Pκ )← W ⊗ G(x, y)dxdy, 1 where the first equality comes from (3.22). It will follow that ← − W∗ (Px,y ⊕ Pκ ) ⊗ G(x, y)dxdy = W∗ (Pκ ) ⊗ M, (3.24) where M = MQ1 is the kernel from W(ΣW ) to Q1 given by M(Φ, U ) = m+ {Φ−1 (U )}. ˆ : ΣW × [0, ∞) → ΣC × Q1 is injective by Lemma 12. For To see this, observe that W E ∈ W(ΣW ) and U ⊂ Q1 , let (Φ, z) ∈ E × U , then there is a unique λ ∈ ΣW and t ∈ [0, ∞) with W(λ) = Φ and Φ(t) = z. Then ˆ ∗ (Pκ ⊗Cκ m+ ){E×U } = Pκ ⊗Cκ m+ {W ˆ −1 (E×U )} = W W −1 (E) m+ {(W(λ))−1 (U )}dPκ (λ) m+ {Φ−1 (U )}dW∗ (Pκ )(Φ) = W∗ (Pκ ) ⊗ M{E × U }. = E Since the measures agree on all cylinder sets, they must be the same. For any U ⊂ Q1 , the kernel MU is the restriction of M to W(ΣW ) × U , and so (3.20) 92 follows from restricting both sides of (3.24) to W(ΣW ) × U . Finally, if we apply the inverse ˆ −1 to (3.20), we get equation (3.19). map W Corollary 1. Suppose U ⊂ [0, ∞) × [0, ∞) with U G−4,−4 (x, y)dxdy < ∞. If the laws of the welding curve of an extended BSLEκ (−4, −4) process started from (0; x, −y) are integrated against IU G−4,−4 (x, y)dxdy, it results in a bounded measure on curves which is absolutely continuous with respect to the law of the BSLEκ welding curve, and the RadonNikodym derivative is |MU (Φ, Q1 )| = Cκ m+ (Φ−1 (U )). Proof. For any measurable E ⊂ W(ΣW ), (3.20) implies that U ← − W∗ (Px,y ⊕ Pκ ){E}G(x, y)dxdy = W∗ (Px,y ⊕ Pκ ) ⊗ IU G(x, y)dxdy{E × Q1 } = W∗ (Pκ ) ⊗ MU {E × Q1 } = = E 3.6 E MU (Φ, Q1 )dW∗ (Pκ ){Φ} Cκ,1 m+ {Φ−1 (U )}dW∗ (Pκ ){Φ}. One-point estimate The goal of this section is to establish a one-point estimate for the BSLE welding, similar to the one-point estimate used to establish the Green’s function for forward SLE. After doing this, we will see that G(x, y) = G−κ−4,−κ−4 (x, y) = x1+4/κ y 1+4/κ (x + y)−4−κ/2−8/κ serves as an appropriate Green’s function to use to condition welding x and y together. This is the only special case in which we have been able to construct Ga,b via a geometrically motivated 93 limit procedure. Fix any x, y > 0. We will again be doing Ito’s formula calculations using the definitions √ from Section 3.3. For clarity, we will review the definitions. Let Xt = ft (x) − κBt and √ Yt = κBt − ft (−y) so that both X and Y are positive. Ito’s formula implies that √ √ 2 2 dXt = − κdBt − dt, and dYt = κdBt − dt. Xt Yt Moreover, −2 d(Xt + Yt ) = dt. Xt + Yt Xt Yt (3.25) X −Y Define Wt = Xt +Yt as before, which is a process which stays in (−1, 1) for as long as it t t exists. It only fails to exist if τx = τ−y , where τz is the time so that z is swallowed by the BSLE traces. If this does not happen (which is Pκ - a.s.), then Wt is constant at ±1 after τ = τx,y := min{τx , τ−y }. What we want to estimate is P(E ), where E = {Xτ + Yτ ≤ }. Theorem 7. For any x, y > 0, with the notation above, there is a constant C < ∞ depending only on κ so that 1 lim − 2 (κ+4) P(E ) = Cx1+4/κ y 1+4/κ (x + y)−4−κ/2−8/κ := G(x, y). →0 Proof. By Ito’s formula, (3.26) √ −2 κ 4Wt dWt = dBt + dt. Xt + Yt Xt Yt Performing the random time change u(t) = 0t κ dr, we get X r Yr ˜t = − 1 − W ˜ 2 dB ˜t + 4 W ˜ t dt, dW t κ 94 (3.27) ˜ t = W −1 . Recall that W ˆ t is a radial Bessel process. where B˜s is a Brownian motion and W u (t) Observe that uτ is the lifetime of a process satisfying (3.7). By [39], there exists a constant C depending only on κ so that for any t > 0, P{uτ > t} ≤ C exp{ By equation (3.5), we have Xt + Yt = 8 −1 (2 + )t}. 2 κ (3.28) κ if and only if ut = − ln 2 x+y := h( ), and E = {uτ > h( )}. Thus, equation (3.28) implies P(E ) = P{uτ > −κ −1 ln ≤ C exp{ (2 + 8/κ) 2 2 −κ ln 2 x+y x+y }=C } 1 (κ+4) 2 x+y . (3.29) In order to find the exact value of equation (3.26), we make use of a more precise bound for P(uτ > t). Equation (B.14) in [39] states that |P{T > t} − 2˜ p(t, w)| ≤ Ce−(6+12/κ)t , (3.30) where w = (x − y)/(x + y) and p˜(t, w) = Cκ (1 − w2 )1+4/κ e−(1+4/κ)t . Inequality (3.29) implies that 3κ+6 |P(E ) − 2˜ p(h( ), w)| ≤ C exp{−(6 + 12/κ)(−κ/2) ln 95 x+y }=C x+y . Multiplying through by −1/2(κ+4) gives | −1/2(κ+4) P(E ) − 2 −1/2(κ+4) p˜(h( ), w)| ≤ C 3κ+6 1 5/2κ+4 → 0 x+y as → 0. Thus, lim −1/2(κ+4) P(E ) = lim 2 −1/2(κ+4) p˜(h( ), w), →0 →0 if this second limit exists. By the definition of p˜, we have p˜(h( ), w) = Cκ (1 − w2 )1+4/κ 1/2(κ+4) x+y , and therefore, using w = (x − y)/(x + y), we get −1/2(κ+4) p˜(h( ), w) = Cκ (1 − w2 )1+4/κ (x + y)1/2(κ+4) for every > 0. 96 = Cκ x1+4/κ y 1+4/κ (x + y)−4−κ/2−8/κ Chapter 4 Measure driven chordal Loewner evolution 4.1 Theorem statement and definitions For any T > 0, let NT denote the set of positive measures µ on [0, T ] × R whose first coordinate marginal is Lebesgue measure m on the Borel subsets of [0, T ]. That is, for any Borel set E ⊂ [0, T ], we have µ(E × R) = m(E). We also assume that the support of µ is compact. Define RT to be the minimal R > 0 so that µ|[0,T ]×R = µ|[0,T ]×[−R,R] . This set of measures has a natural topology induced by weak convergence of measures. That is, µn → µ in NT if for every bounded φ ∈ C([0, T ] × R), where C(X) = continuous functions on X, we have lim n→∞ [0,T ]×R φdµn → φdµ. [0,T ]×R Similarly, define the space N by µ ∈ N if µ is a measure on [0, ∞) × R which has locally compact support and whose first coordinate marginal is Lebesgue measure. Then we define the natural projections PT : N → NT by PT (µ) = µ|[0,T ]×R . The topology on N is determined by µn → µ in N if PT (µn ) → PT (µ) for every positive T . Let G be the set of familes of conformal maps (gt ) with gt : H\Kt → H, where (Kt ) is a family of growing H-hulls with hcap(Kt ) = 2t for every t. We will always use (ft ) to 97 represent the family of inverse maps ft : H → H\Kt . We put a topology on G induced by (gtn ) → (gt ) if ftn (z) → ft (z) locally uniformly in (t, z). For µ ∈ N , we say that (gt ) solves the chordal Loewner equation driven by µ if gt (z) − z = Φ(u, gs (z))dµ(s, u), [0,t]×R Φ(u, z) = 2 , z−u (4.1) for every z ∈ H\Kt , where Kt is the set where the solution blows up by time t. Theorem 8. a) There is a one-to-one correspondence between measures µ ∈ N and con- tinuously growing families of H-hulls (Kt ) with hcap(Kt ) = 2t. This correspondence is given by the unique solution (gt ) of (4.1) driven by µ, where gt = gKt . b) Suppose {µn } and µ in N have solutions to (4.1) given by (gtn ) and (gt ) respectively. Assume further that for all t > 0, the associated hulls {Ktn }n∈N are uniformly bounded. Then µn → µ in N if and only if the associated hulls Ktn → Kt in the Carath´eodory topology. The basic argument for the proof is as follows. First, given a measure µ ∈ N , we use the general form of the chordal Loewner equation in [7] to approximate a disintegration n of µ as dµn t dt → dµt dt = dµ. For each n, the family of measures dµt define a family of chordal Loewner hulls (Ktn )t>0 with corresponding Loewner maps (gtn )t>0 . We show that subsequences of these converge to a family of conformal maps (gt ) and chordal hulls (Kt ) such that gt solves (4.1). To show the correspondence in the other direction, we start with a family of hulls (Kt ). For any hull K with hcap(K) = 2T we show that there is some measure µT so that µT can be associated with K. For the whole process (Kt ), we use this construction to build measures µδn which account for growth in each interval of length δ. 98 Taking the limit of this approximation as δ → 0, we get a measure µ which corresponds to the family of hulls (Kt ). 4.2 Preliminary results Theorem 8 mentions convergence in the Carath´eodory sense. The definition of this convergence can be found in [26], and will be restated here. Let {Dn }n∈N and D be domains in C. We say that Dn → D in the Carath´eodory sense if (i) For any compact K ⊂ D, there is an N so that K ⊂ Dn for all n ≥ N . (ii) For any z ∈ ∂D, there exists a sequence zn ∈ ∂Dn so that zn → z. Hulls K n are said to converge in the Carath´eodory sense to K if their complements converge. That is, if H\K n → H\K in the Carath´eodory sense. The following lemma can be found in [26]. Lemma 17. Suppose that Dn → D as domains, gn : Dn → C are conformal, and g : D → C is a function so that gn → g locally uniformly on D. Then either g is constant or g is conformal. In the latter case, the following statements hold: (a) gn (Dn ) → g(D) in the Carath´eodory topology, (b) gn−1 → g −1 locally uniformly in g(D). Here we provide the proof that the assumptions of the classical existence and uniqueness theorem are satisfied for the chordal Loewner ODE driven by a weakly continuous family of measures on R as introduced in [7]. 99 Proposition 7. Suppose that {νt }t≤T is a family of positive Borel measures on R so that t → νt is continuous in the weak topology. Further, assume that for all t, there exists some Mt > 0 so that νs (R) ≤ Mt and supp{νs } ⊂ [−Mt , Mt ] for all s ≤ t. Then for all z ∈ H, there exists a unique solution to the differential equation ∂t gt (z) = 2 dνt (z), R gt (z) − u g0 (z) = z on some maximal interval [0, Tz ). Proof. Fix z ∈ H and T > 0. Define R = {(t, w) : 0 ≤ t ≤ T, Im(w) ≥ Im(z)/2}. Let ψ(t, w) = 2 dνt (u), R w−u so that ∂t gt (z) = ψ(t, gt (z)). By the classical local and maximal existence and uniqueness theorems for ODEs, if both ψ and ∂w ψ are continuous in some [0, t] × B(z, r) ⊂ R, then there exists a solution to the ODE in some maximal time interval [0, Tz ). First, we compute ∂w ψ(t, w). Observe that if w, ξ have imaginary parts at least Im(z)/2, then ψ(t, w) − ψ(t, ξ) 1 = w−ξ R (w − ξ) = 2 2 − w−u ξ−u dνt (u) 2 ξ−w −2 · dνt (u) → dν (u), 2 t R w − ξ (w − u)(ξ − u) R (w − u) where the limit is as ξ → w. This follows by the dominated convergence theorem, since |ξ − u|, |w − u| ≥Im(z)/2 and νt is a finite measure. Next, we will prove that ψ is continuous in the desired region. Suppose that (tn , wn ) → (t, w), with (tn , wn ) ∈ R. Then 100 |ψ(tn , wn ) − ψ(t, w)| ≤ 2 2 dνtn (u) − dνtn (u) + R wn − u R w−u Since Im(w) ≥Im(z)/2, the function u → 2 2 dνtn (u) − dνt (u) . R w−u R w−u 2 is uniformly bounded, so weak continuity w−u implies that the second term approaches 0. To estimate the first term, the map w → (w − u)−1 is Lipschitz with constant C depending only on the fixed z, so 2 2 dνtn (u) − dνtn (u) ≤ C |wn − w|dνtn (u) ≤ CMT |wn − w| → 0. R wn − u R w−u R Thus, ψ is continuous in the desired region. The proof for ∂w ψ will be identical based on the above formula, only it will use the Lipschitz constant for w → (w − u)−2 . We need to establish an important result about H-hulls. In [36], it is proven that there ∗ so that µ is a positive measure µK with support in SK K =hcap(K), and fK (z) − z = −1 dµ (u). ∗ z−u K SK Lemma 18. If K is an H-hull, then for all z ∈ H, |fK (z) − z| ≤ 7diam(K). Proof. Suppose without loss of generality that K ⊂ Kr := {z ∈ H : |z| ≤ r}, where ∗ ⊂ S ∗ = [−2r, 2r]. r = max{|z| : z ∈ K}. In [7], it is proven that hcap(Kr ) = r2 and SK Kr 101 Fix any L > 2r, and suppose that |z| ≥ L. Then |fK (z) − z| ≤ µK r2 1 dµK (u) ≤ ≤ . ∗ z−u L − 2r L − 2r SK Suppose that |z| ≤ L. Then the maximum modulus principle implies that |fK (z)| ≤ sup {|w|=L}∪{−L0,n∈N on R so that (a) Each µn t is a probability measure on R. (b) For every n ∈ N, the map t → µn t is continuous with respect to the weak topology on measures on R. 102 n (c) The measure given by dµn t dt is in N , and dµt dt → µ as n → ∞ in N . (d) For almost every t ∈ [0, ∞), there is a probability measure µt so that µn t → µt weakly. Also, dµt dt = dµ. Proof. First, we must define the measures µn t . For each n ∈ N, t > 0, define a linear functional Ln,t : Cc (R) → C by Ln,t (φ) = n φ(u)dµ(s, u). (4.2) [t,t+1/n]×R To see that this map is well defined, observe that |Ln,t (φ)| ≤ n [t,t+1/n]×R |φ(u)|dµ(s, u) ≤ φ ∞ nµ([t, t + 1/n] × R) = φ ∞ < ∞. The map Ln,t is clearly a positive linear functional. By the Riesz representation theorem, there exists a Radon measure µn t defined on the Borel sets of R so that equation (4.2)= n R φdµt . To prove (a), for every k ∈ N, define φk ∈ Cc (R) so that φk (u) = 1 for u ∈ [−k, k], and φk (u) → 1 monotonically for every u. By the monotone convergence theorem (applied twice) and equation (4.2), µn t (R) = lim k→∞ R φk dµn t = lim n k→∞ [t,t+1/n]×R φk (u)dµ(s, u) = nµ([t, t + 1/n] × R) = 1. We will now make several arguments about the measures converging weakly on R. For weak convergence, we need to test all bounded continuous functions on R. However, we 103 are testing the local properties of the measures µ before some time T . We assume that our measure µ, when restricted to [0, T ] × R, has support contained in [0, T ] × [−RT , RT ]. Therefore, if φ is any bounded continuous function, there exists some φ∗ ∈ Cc (R) so that φ∗ = φ on [−RT , RT ]. Hence, for all weak convergence arguments below, we will make this reduction automatically and work only with continuous functions with compact support. To see (b), suppose that tk → t in [0, ∞). Then tk → t in [0, T ] for some T . Let φ ∈ Cc (R) be arbitrary. Then φdµn t =n k R =n φ(u)dµ(s, u) [tk ,tk +1/n]×R I[t ,t +1/n]×R (s, u)φ(u)dµ(s, u). k k [0,T ]×R (4.3) Then I[t ,t +1/n]×R (s, u)φ(u) → I{[t,t+1/n]×R} (s, u)φ(u) µ-almost everywhere, and is domik k nated by φ ∞ ∈ L1 ([0, T ]×R, µ). Thus, Lebesgue’s dominated convergence theorem implies that lim (4.3) = n k→∞ [0,T ]×R φdµn t. I{[t,t+1/n]×R} (s, u)φ(u)dµ(s, u) = R Now we must show c), starting with dµn t dt ∈ N . For any Borel set E ⊂ R, part (a) implies that E 1dµn t dt = R dt = m(E). E Also, if t ≤ T , then it is clear that the support of µn t is contained in [−RT +1 , RT +1 ]. Therefore, dµn t dt ∈ N . To show convergence, suppose that φ ∈ Cc ([0, T ] × R). For any T > 0, [0,T ]×R φ(t, u)dµn t (u)dt = [0,T ] φ(t, u)dµn t (u) dt R T = n 0 φ(t, u)dµ(s, u) dt [t,t+1/n]×R 104 s = s T + [0,1/n]×R 0 + [1/n,T ]×R s−1/n nφ(t, u)dtdµ(s, u). [T,T +1/n]×R s−1/n Since φ has compact support, fixing n guarantees that this function is L1 (µ × m), so the above use of Fubini’s theorem is justified. We will show that the first and third terms go to 0, and the middle term converges to [0,T ]×R φ(s, u)dµ(s, u). To bound the first term, observe that s s nφ(t, u)dtdµ(s, u) ≤ φ n 1dtdµ(s, u) [0,1/n]×R 0 [0,1/n]×R 0 sdµ(s, u) ≤ φ n = φ n [0,1/n]×R 1 φ dµ(s, u) = , n [0,1/n]×R n which converges to 0 as n goes to infinity. The third integral converges to 0 by a similar argument. Thus, it suffices to show that lim n→∞ [1/n,T ]×R s 1 φ(t, u)dt dµ(s, u) = φ(s, u)dµ(s, u). 1/n s−1/n [0,T ]×R (4.4) For each u, the map t → φ(t, u) is continuous on [0, ∞), so the fundamental theorem of calculus implies that s 1 φ(t, u)dt = φ(s, u) n→∞ 1/n s−1/n lim for each s. Thus, for µ-almost every (s, u), s I{[1/n,T ]×R} (s, u) n φ(t, u)dt s−1/n 105 → I{[0,T ]×R} (s, u)φ(s, u) as n → ∞. Also, for every n ∈ N, s > 0, and u ∈ R, we have s I{[1/n,T ]×R} (s, u)n s−1/n φ(t, u)dt ≤ φ ∞ I[0,T ]×[−N,N ] (s, u) ∈ L1 (µ), where the support of φ is contained in [0, T ]×[−N, N ] for some N > 0. Therefore, Lebesgue’s dominated convergence theorem implies that (4.4) holds, which completes the proof of (c). We now prove (d). Since µn t is a collection of Radon probability measures with uniformly compact support for each t, there is a subsequential weak limit µt . Let {φk }k∈N be a dense countable subset of Cc (R), which is separable. We claim that for each k, there exists a subset Ek ⊂ [0, ∞) so that m(Ekc ) = 0, and for all t ∈ Ek we have lim φ (u)dµn t (u) n→∞ R k = φk (u)dµt (u). R To see this, define a measure νk on R by νk (E) := E×R φk (u)dµ(s, u), which is absolutely continuous with respect to Lebesgue measure. By ([28], Theorem 7.14) we have for m-almost all t (where a.e. depends on φk , hence the set Ek ), φk dµn t = R Thus, limn φk dµn t → ν [t, t + 1/n] dν 1 φk (u)dµ(s, u) = k → k (t). 1/n [t,t+1/n] m[t, t + 1/n] dm φk dµn t exists, and there is subsequential limit equal to φk dµt , and therefore φk dµt for each k. Let E = ∩k Ek , so that m(E c ) = 0, and suppose t ∈ E. Let φ ∈ Cc (R). We claim that φdµn t → φdµt as n → ∞. Let > 0. Then there exists some k so that φk − φ ∞ < . 106 Then for sufficiently large n, φ(u)dµn t (u) − φ − φk dµn t + ≤ R φ(u)dµt (u) R R φk dµn t − R φk − φdµt ≤ + + . φk dµt + R R Thus, µn t → µt weakly for all t ∈ E, and hence for m-almost every t. The final claim is true because for any φ ∈ Cc ([0, T ] × R), for any u and almost any s, we have R φ(s, u)dµs (u) = limn→∞ R φ(s, u)dµn s (u). Therefore, the dominated convergence theorem implies that t t φ(s, u)dµs (u)ds = 0 R t φ(s, u)dµn s (u)ds = = lim n→∞ 0 lim 0 n→∞ R R φ(s, u)dµn s (u)ds φ(s, u)dµ(s, u). [0,t]×R The last equality follows from part (c), since dµn t dt → dµ in N . Remark: For each t > 0 so that µt as in part (d) exists, we can exactly repeat this construction with the functional φ→n φ(u)dµ(s, u) [t−1/n,t]×R for each n. The resulting measures will converge to the same µt for almost all t. Therefore, 107 if h < 0, we can use the convention φ(s, u)dµ(s, u) = − [t,t+h]×R φ(s, u)dµ(s, u) [t+h,t]×R as in Riemann integration to discuss differences gt+h (z) − gt (z) without making reference to whether h is positive or negative. 4.4 Elementary calculations Next, we will make several calculations assuming that (4.1) driven by µ has a solution (gt ). These calculations will then be applied to the measures dµn = dµn t dt, which have solutions (gtn ) for each n by Proposition 7. Moreover, the maps gtn are all conformal, and a solution to (4.1) will be constructed as a locally uniform limit of gtn , and so we can treat any solution as conformal. We state and prove these lemmas in their more general state since we will need them after a solution is proven to exist. First, we have to prove two Lipschitz estimates which will be used repeatedly. Lemma 20. a) If z, w ∈ H with imaginary parts at least h > 0 and u ∈ R, then 2 |Φ(u, z) − Φ(u, w)| ≤ 2 |z − w|. h b) Suppose (gt ) is a solution to the chordal Loewner equation driven by µ ∈ N . Then for any z ∈ H and t1 < t2 ≤ T < τz , |gt2 (z) − gt1 (z)| ≤ 108 2 |t − t1 |. Im(gT (z)) 2 Proof. To prove a), observe that |z − u| ≥ h and |w − u| ≥ h. Therefore, z−w 2 2 2 − =2 ≤ 2 |z − w|. z−u w−u (z − u)(w − u) h |Φ(u, z) − Φ(u, w)| = To prove b), observe that (4.1) implies that the imaginary part of gt (z) is a decreasing function of t. Therefore, |gt2 (z) − gt1 (z)| = ≤ 2 2 dµ(s, u) ≤ dµ(s, u) [t1 ,t2 ]×R gs (z) − u [t1 ,t2 ]×R |gs (z) − u| 2 2 2 dµ(s, u) ≤ dµ(s, u) = |t − t1 |. Im(gT (z)) 2 [t1 ,t2 ]×R Im(gs (z)) [t1 ,t2 ]×R Im(gT (z)) Lemma 21. If (gt ) is a solution to the chordal Loewner equation driven by µ ∈ N and z ∈ H. Then for almost any t < τz , ∂t gt (z) = 2 dµt (u), R gt (z) − u where µt is the measure from Lemma 19 part d). Proof. This follows from Lemma 19 part d), which says that dµt (u)dt = dµ(t, u). In particular, if T < τz , Lemma 19 part d) implies that for t < T , t gt (z) − z = Φ(u, gs (z))dµ(s, u) = [0,t]×R 0 2 dµs (u)ds. R gs (z) − u Since 2 2 dµs (u) ≤ ∈ L1 ([0, T ]), g (z) − u Im(g (z)) R s T 109 2 the fundamental theorem of calculus implies that ∂t gt (z) = R g (z)−z dµt (z) for almost every t t. We can now prove uniqueness. Proposition 8. A solution of the chordal Loewner equation driven by µ ∈ N must be unique. Proof. Fix z ∈ H, and suppose that gt and g˜t are two solutions of the integral equation gt (z) = z + 2 dµ(s, u) [0,t]×R gs (z) − u up to some time T + . Then {Im(gt )}t≤T is decreasing and bounded below by Im(gT ). The same is true for g˜t , so we can assume that there is a constant C so that for all t ≤ T , we have Im(gt ) and Im(˜ gt ) are bounded below by C. By Lemma 21, for almost all t ≤ T (depending on z, which is fixed), we know that ∂t (gt − g˜t ) = 2 g˜t − gt dµt (u). gt − u) R (gt − u)(˜ Therefore, |∂t (gt − g˜t )| ≤ 2 |˜ gt − gt | 2 2 dµt ≤ 2 |˜ gt − gt |dµt (u) = 2 |gt − g˜t |. gt − u)| C R C R |(gt − u)(˜ Since gt and g˜t are absolutely continuous by Lemma 20 , we get that t |gt − g˜t | ≤ |∂s (gs − g˜s )|ds ≤ 0 t 2 |gs − g˜s |ds. C2 0 By Gronwall’s inequality applied to |gt − g˜t |, it follows that gt = g˜t for all t ≤ T . Since T was an arbitrary number below the lower blow-up time, it follows that gt = g˜t for all t. 110 Next, we need to derive some basic continuity and differentiability properties of the inverse function f = ft (z) = gt−1 (z). Lemma 22. Suppose that g(t, z) = gt (z) solves the chordal Loewner equation driven by µ ∈ N , and let ft (z) = gt (z)−1 . Then both f and f are continuous in the variables (t, z). Proof. The main part of this lemma is that if (Kt ) is an increasing family of hulls with hcap(Kt ) = 2t, then the map t → H\Kt is continuous in the Carath´eodory topology. Then Lemma 17, part (b) implies that if tn → t, then ftn → ft locally uniformly in H\Kt , and therefore ftn → ft locally uniformly in H\Kt . It suffices to prove that H\Ktn → H\Kt in the cases where tn increases to t and tn decreases to t separately. First, assume that tn decreases to t. Then Kt ⊂ Ktn for all n. I claim that ∩n Ktn = Kt . Let K = ∩n Ktn , which is an H-hull with Kt ⊂ K , and assume that K = Kt . Then hcap(K ) >hcap(Kt ). However, 2tn =hcap(Kn ) ≥hcap(K ) >hcap(Kt ) = 2t, which contradicts the assumption that tn → t. Thus, Kt = ∩n Ktn . The properties for Carath´eodory convergence are readily verified Now, suppose that tn increases to t. Note that we do not necessarily have ∪n Ktn = Kt , so we must argue differently. Observe that H\Kt ⊂ H\Ktn for each n, so gtn is defined on H\Ktn for each n. Let Dn = D = H\Kt . Then Dn → D clearly. Also, gtn → gt locally uniformly, and gt is not constant. By Lemma 17 (a), we have that gtn (H\Kt ) → H. Applying it again to the inverse maps yields H\Ktn → H\Kt . n Lemma 23. Let µ ∈ N , and let dµn = dµn t (u)dt, where µt is as in Lemma 19. Then there 111 are functions (gtn ) which solve (4.1) driven by µn . If ftn = (gtn )−1 , then ftn (z) − z = − 2(fsn ) (z) n dµ (s, u) [0,t]×R z − u (4.5) for each z ∈ H, t > 0, n ∈ N. Proof. For each n, Lemma 19 part b) and Proposition 7 imply that there is a solution of ∂t gtn (z) = 2 dµn t (z), n g (z) − u R t g0n (z) = z, which is the solution of (4.1) driven by µn . This equation holds for all t < τz . If ftn = (gtn )−1 , the chain rule then implies that ∂t ftn (z) = −2(ftn ) (z) n dµt (u), z−u R f0n (z) = z. Integrating this gives with respect to t yields (4.5). We are now almost ready to prove existence of a solution to the chordal measure driven Loewner equation. In order to use a limiting argument, we need some sort of control on the size of the hulls {Ktn }t≤t0 for any fixed t0 . This control will be provided by the following two lemmas. Lemma 24. Suppose (gt ) is a solution of the chordal Loewner equation driven by µ ∈ N . √ For each t, define Mt = max{ t, Rt }. Then for all z ∈ Kt , we have |z| ≤ 4Mt . Moreover, if |z| > 4Mt , then |gs (z) − z| ≤ Mt for all 0 ≤ s ≤ t. Remark: This lemma is an extension of Lemma 4.13 in [7], where it is proven for the 112 chordal Loewner equation driven by a continuous function. The proof is different, however. The argument follows the standard proof for the existence of a solution of an ODE. Proof. Fix t > 0 and z ∈ H with |z| ≥ 4Mt . Recall that Rt is defined to be Rt = inf{R > 0 : µ|[0,t]×R is supported in [0, t] × [−R, R]}. We will use a Picard iteration argument to show that τz ≥ t, where τz is the lifetime of (4.1) at z. Let w ∈ B(z, Mt ), which implies |w| ≥ 3Mt . For any u ∈ [−Rt , Rt ], we have |w − u| ≥ |w| − |u| ≥ 3Mt − Mt = 2Mt . Therefore, if Φ(u, z) = 2 , we have for any w, w ∈ B(z, Mt ), z−u |Φ(u, w) − Φ(u, w )| = 2|w − w | 1 2(w − w ) ≤ = |w − w |. (w − u)(w − u) 4Mt2 2Mt2 This implies that Φ(u, w) has Lipschitz constant 1 for w ∈ B(z, Mt ). Also, if w ∈ 2Mt2 B(z, Mt ) and u ∈ [−Rt , Rt ], |Φ(u, w)| ≤ 2 1 = . 2Mt Mt Then the restriction of Φ to [−Rt , Rt ] × B(z, Mt ) satisfies Φ ∞ ≤ 1/Mt . To perform the Picard iteration, fix an a > 0 to be determined later, and define an operator Λ on the set of continuous functions from [0, a] to B(z, Mt ) by Λ(φ)(r) = z + Φ(u, φ(s))dµ(s, u), 0 ≤ r ≤ a. [0,r]×R We will show that for each a < t, Λ(φ) will be a continuous map from [0, a] to B(z, Mt ). If a < t, then a < Mt2 by construction of Mt . For r ≤ a, |Λ(φ)(r) − z| = Φ(u, φ(s))dµ(s, u) [0,r]×R 113 |Φ(u, φ(s))|dµ(s, u) ≤ ≤ [0,r]×R a ≤ Mt . Mt (4.6) Therefore, Λ(φ) sends [0, a] into B(z, Mt ). Next, we need to show that Λ is a contraction. For any continuous φ1 , φ2 : [0, a] → B(z, Mt ), the Lipschitz estimate implies that for any r ∈ [0, a], |Λ(φ1 )(r) − Λ(φ2 )(r)| ≤ ≤ [0,r]×R |Φ(u, φ1 (s)) − Φ(u, φ2 (s))| dµ(s, u) 1 a φ1 − φ2 ∞ dµ(s, u) ≤ 2 φ1 − φ2 ∞ , 2 Mt [0,a]×R 2Mt where a/Mt2 < 1. Thus, Λ is a contraction between Banach spaces, and the Banach fixed point theorem implies that there is a unique continuous φ : [0, a] → B(z, Mt ) so that φ(r) = Λ(φ)(r) = z + 2 dµ(s, u), [0,r]×R φ(s) − u 0 ≤ r ≤ a. By Proposition 8, φ(r) = gr (z) for all r ≤ a < t. Thus, τz ≥ t. Moreover, if |z| > 4Mt , this construction can be stretched to a = t, and so applying (4.6) to gt (z) proves |gs (z) − z| ≤ Mt for all s ≤ t. n Lemma 25. If µ ∈ N , let µn t be as in Lemma 19. For each n, let (gt ) be the Loewner maps n driven by {µn t }t≤T with corresponding H-hulls (Kt )t≤T . Then for every t0 > 0, there exists some Mt0 > 0 so that Ktn ⊂ {z ∈ H : |z| ≤ Mt0 }. 0 Proof. By Lemma 24, it suffices to show that the support of the measures {µn t }t≤t0 ,n∈N are uniformly bounded. We have assumed that µ restricted to [0, t0 + 1] × R is supported in [0, t0 + 1] × [−R, R] for some R > 0. Suppose that x ∈ R and > 0 so that (x − 2 , x + 2 ) × 114 [t, t + 1/n) ⊂ ([0, t0 + 1] × [−R, R])c . In particular, µ((x − , x + ) × [t, t + 1/n]) = 0. Let φk ∈ Cc (R) be a monotonic sequence converging to I(x− ,x+ ) . Then µn t (x − , x + ) = lim k→∞ R φk (u)dµn t (u) = lim n k→∞ [t,t+1/n]×R φk (u)dµ(s, u) = nµ((x − , x + ) × [t, t + 1/n]) = 0. n Since the supports of the measures µn t are uniformly bounded, it follows that the hulls Kt are bounded independently of n. 4.5 Proof of existence Using the boundedness of hulls from Lemma 25, we can prove equicontinuity of some subsequence of any family f n of inverse functions associated with a sequence of measures µn ∈ N associated with µ. Lemma 26. Let µ ∈ N , and suppose (gtn )t≥0 is the family of conformal maps which solve the chordal Loewner equations driven by the measures dµn (t, u) = dµn t (u)dt for each n. Then there exists a family of conformal maps (gt ) so that a subsequence of (gtn ) converges to (gt ) in G. Remark: The proof of this lemma will hold for any µn ∈ N with µn → N as long as (4.5) holds at each n, and if the hulls (Ktn ) are uniformly bounded for each t. We will do this to prove that Theorem 8 part b) holds. Proof. Fix T < ∞. Applying Lemma 18 to each member of the family (Ktn )t≤T,n∈N , and us115 ing the assumption that the hulls are uniformly bounded, we get that the family (ftn )t≤T,n∈N is normal in H. We will use this normality and equation (4.5) to show that the functions f n are equicontinuous in both t and z on any compact set [0, N ] × RN ⊂ [0, ∞) × H, where RN = [−N, N ] × [1/N, N ]. Since (ftn )t≤T,n∈N is a normal family on H, so are the families of derivatives (ftn )t≤N,n∈N and (ftn )t≤N,n∈N . Since RN ⊂ H is compact, this implies that there are constants CN , CN < ∞ with |(fsn ) (z)| ≤ CN and |(fsn ) (z)| ≤ CN for all s ≤ N and z ∈ RN . Therefore, |(fsn ) (z1 ) − (fsn ) (z2 )| ≤ CN |z1 − z2 |, for all s ≤ N, z ∈ RN . (4.7) If 0 ≤ t1 ≤ t2 ≤ N and z1 , z2 ∈ RN , equation (4.5) implies that |ftn (z1 ) − ftn (z2 )| ≤ |z1 − z2 | + 2 2 1 +2 (fsn ) (z1 ) (fsn ) (z2 ) n − dµ (s, u) z2 − u [0,t1 ]×R z1 − u (fsn ) (z2 ) n dµ (s, u) = |z1 − z2 | + 2(I) + 2(II). [t1 ,t2 ]×R z2 − u If we can bound (I) and (II) uniformly in N , |t2 − t1 |, and |z1 − z2 |, we will have that f n is equicontinuous in [0, N ] × RN . For any u ∈ R and z ∈ RN , we have |z − u| ≥ 1/N . Therefore, (II) can be bounded by (II) ≤ CN 2 dµ(s, u) ≤ CN (2N )|t2 − t1 |. [t1 ,t2 ]×R |z − u| To estimate (I), observe that for any u ∈ R, we have (fsn ) (z1 ) (fsn ) (z2 ) (fsn ) (z1 ) (fsn ) (z1 ) (fsn ) (z1 ) (fsn ) (z2 ) − ≤ − + − z1 − u z2 − u z1 − u z2 − u z2 − u z2 − u 116 ≤ CN z2 − z1 + N (fsn ) (z1 ) − (fsn ) (z2 ) (z1 − u)(z2 − u) ≤ N 2 CN |z1 − z2 | + N CN |z1 − z2 |, where the last inequality follows from (4.7). Therefore, since the marginal of µn is Lebesgue measure and t1 ≤ N , we have (I) ≤ [0,t1 ]×R N 2 CN |z1 − z2 | + N CN |z1 − z2 |dµ(s, u) ≤ t1 N 2 CN + N CN |z1 − z2 | ≤ N 3 CN + N 2 CN |z1 − z2 |. Therefore, we have equicontinuity of f n on [0, N ] × KN for every N . By the ArzelaAscoli theorem, there is a subsequence so that f nk → f N for each N locally uniformly on [0, N ] × KN . Taking N → ∞ and applying a diagonal argument gives the desired limit function f : [0, ∞) × H. Lemma 27. If (gtn ) is a solution of the chordal Loewner equation driven by µn , (gtn ) → (gt ) in G, and µn → µ in N , then (gt ) is a solution to the chordal Loewner equation driven by µ. Proof. The chordal equation (4.1) implies that gtn (z) = [0,t]×R Φ(u, gsn (z)) − Φ(u, gs (z))dµn (s, u) + 117 Φ(u, gs (z))dµn (s, u) + z. (4.8) [0,t]×R To estimate this, we will need to show that lim sup {|gsn (z) − gs (z)|} = 0. n→∞ 0≤s≤t (4.9) This will be justified at the end of the proof. It is easy to show that Im(gsn (z)) is decreasing in s for each n, so taking the limit as n → ∞ shows that Im(gs (z)) is also a decreasing function of s. It also follows that Im(gsn (z)) ≥ min{Im(gt (z)), Im(gt1 (z)), · · · } = h > 0 for each n ∈ N and s ≤ t. Strict inequality follows since {gtn (z)}∞ n=1 is a sequence of numbers in H converging to gt (z) ∈ H. By Lemma 20, if L = 2/h2 , [0,t]×R |Φ(u, gsn (z)) − Φ(u, gs (z))|dµn (s, u) ≤ [0,t]×R L|gsn (z) − gs (z)|dµn (s, u) ≤ L sup{|gsn (z) − gs (z)|}µ ([0, t] × R) = Lt sup{|gsn (z) − gs (z)|} → 0 s≤t s≤t as n → ∞ by (4.9). To estimate the second term in equation (4.8), observe that for a fixed z, since s → gs (z) is continuous, the mapping (s, u) → Φ(u, gs (z)) is a continuous function. Since µn → µ weakly on [0, t] × R, it follows that Φ(u, gs (z))dµn (s, u) → [0,t]×R Φ(u, gs (z))dµ(s, u). [0,t]×R Therefore, the above and (4.9) imply that gt (z) = lim gtn (z) = lim equation (4.8) = n→∞ n→∞ 118 Φ(u, gs (z))dµ(s, u) + z. [0,t]×R Now we prove (4.9). Suppose otherwise, so that there is some > 0, some subsequence nk , and some sequence sk ∈ [0, t] with n |gskk (z) − gsk (z)| ≥ . (4.10) Passing to a new subsequence if necessary, we can assume that there is some s∗ ∈ [0, t] with sk → s∗ . Since gt (z) is continuous in t, gsk (z) converges to gs∗ (z) as k → ∞. We also claim n that gskk (z) is bounded. To see this, if h is defined as above, 2 dµnk (s, u) nk [0,sk ]×R |gs (z) − u| n |gskk (z) − z| ≤ ≤ 2t 2sk ≤ . h h n Therefore, passing to a subsequence if necessary, we can assume that gskk (z) converges to n some w ∈ H as k → ∞. Since ft k (z) converges to ft (z) locally uniformly in (t, z), it follows n n that z = fskk (gskk (z)) → fs∗ (w), and so gs∗ (z) = w. Since gsk (z) → gs∗ (z), we get that n lim gskk (z) − gsk (z) = gs∗ (z) − gs∗ (z) = 0, k→∞ which contradicts (4.10). Finally, we can prove existence. Proposition 9. a) If µ ∈ N , then the solution to the chordal Loewner equation driven by µ exists. b) If (gt ) is the solution of the chordal Loewner equation driven by µ ∈ N , and if ft = gt−1 , 119 then (4.5) holds for f . That is, ft (z) − z = −2fs (z) dµ(s, u). [0,t]×R z − u c) Assume µn → µ in N , and suppose the respective corresponding Loewner maps are (gtn ) and (gt ). Assume further that the associated hulls {Ktn }n∈N are uniformly bounded for each t > 0. Then gtn → gt in G. n Proof. Lemma 26 can be applied to the sequence dµn t dt ∈ N to conclude that (gt ) → (gt ) subsequentially in G to some family of functions (gt ). By Lemma 27 it follows that (gt ) solves the chordal Loewner equation driven by µ, which proves a). To prove b), Lemma 26 can be applied to f n (after passing to a subsequence if necessary), to conclude that |f n (z) − f (z)| → 0. Therefore, it suffices to show that the difference of (4.5) for ftn (z) and ft (z) goes to 0. To see this, observe that −2fs (z) −2(fsn ) (z) n dµ (s, u) − dµ(s, u) z−u [0,t]×R [0,t]×R z − u ≤2 fs (z) (fsn ) (z) − dµn (s, u)+2 z−u [0,t]×R z − u fs (z) fs (z) n dµ(s, u) − dµ (s, u) . [0,t]×R z − u [0,t]×R z − u The first term goes to 0 by the dominated convergence theorem, since (fsn )s≤t is a normal family, so is the family of derivatives. Therefore, |(f n )s (z)| can be bounded uniformly in s ≤ t. Since fsn (z) → fs (z) locally uniformly in z for all s ≤ t, and so does (f n )s . The second term goes to 0 because µn → µ. The last assertion follows from Lemmas 25, the proof of 26 with b), and 27. In particular, by b) and Lemma 25, the remark after Lemma 26 implies that there is a subsequential limit 120 gt for which f n converges to f uniformly in both space and time. Lemma 27 implies that gt is a solution to the Loewner equation driven by µ. Also, Lemma 27 implies that any subsequential limit of gtn solves the Loewner equation driven by µ, which must be unique by Proposition 8, and therefore the sequence (gtn ) converges to (gt ). 4.6 Growing Hulls In this section, we prove the reverse direction of Theorem 8. We start with a continuously growing family of hulls (Kt ) and construct a measure µ ∈ N whose solution to (4.1) generates the family (Kt ). Proposition 10. Let Kt be growing hulls with hcapKt = 2t for every t. Then there is a measure µ ∈ N so that gt : H\Kt → H, where gt solves the Loewner equation driven by µ. We first prove the following lemma. The proof is identical to that of Lemma 6.5 in [19]. Lemma 28. Let K be an H-hull, and let 2T =hcap(K). Then there exists some µ ∈ NT so that if gt is the solution to the Loewner equation driven by µ, then gT : H\K → H. Proof. Let > 0, and let γ : [0, T ] → H be a simple curve so that γ begins in R, encloses ˆ less than , where K ˆ is the union of K and the convex K, and has Hausdorff distance from K hull of K ∩ R. By ([7], Proposition 4.4), there exists a continuous function λ so that if gt is the solution of the chordal Loewner equation driven by λ , then gt : H\γ [0, t] → H for all t. If we define µ = dδλ (t) dt and take a subsequential weak limit as → 0, Proposition 9 implies that there is a measure µ ∈ N whose solution gt satisfies gT : H\K → H. 121 Proof of Proposition 10. Suppose for now that gt = gKt for each t. Let δ > 0, and let l ∈ {0, 1, 2, · · · }. Then hcap(Kδ(l+1) /Kδl ) = 2δ. By Lemma 28, there exists a measure δ,l µ ˜δ,l on [0, δ] × R so that if g˜t δ,l solves the Loewner equation driven by µ ˜δ,l , then g˜δ : H\(Kδ(l+1) /Kδl ) → H. Define a new measure µδ,l on [δl, δ(l + 1)] × R by shifting the previous measure. That is, µδ,l (E) = µ ˜δ,l ({(x − δl, y) : (x, y) ∈ E}). ∞ δ,l l=0 µ , Finally, define a measure µδ ∈ N by µδ = and let gtδ solve the Loewner equation driven by µδ with gtδ : H\Ktδ → H. δ for every l and δ. We prove this by induction. The case We will show that Kδl = Kδl l = 0 is obvious, since if t ≤ δ, then gtδ (z) = z + [0,t]×R Φ(u, gsδ (z))d˜ µδ,0 (s, u), δ,0 and therefore gδδ = g˜t (z) for each t ≤ δ, and Kδ /K0 = Kδ . δ : H → H\K . For any t ∈ [0, δ], Suppose this is true for l. Then fδl δl δ,l δ δ (z)) = f δ (z) + Gt (z) := gδl+t (fδl δl + [δl,δl+t]×R δ (z)))dµδ (s, u) Φ(u, gsδ (fδl δ (z) + (g δ (f δ (z)) − f δ (z)) + = fδl δl δl δl =z+ [0,t]×R [0,δl]×R δ (z)))dµδ (s, u) Φ(u, gsδ (fδl [δl,δl+t] δ (z)))dµδ,l (s, u) Φ(u, gsδ (fδl δ δ (z)))d˜ Φ(u, gδl+s (fδl µδ,l (s, u) 122 δ,l =z+ [0,t]×R Φ(u, Gs (z))d˜ µδ,l (s, u). δ,l δ,l Thus, Gt solves the Loewner equation driven by µ ˜δ,l , and so Gδ = g˜δδ : H\(Kδ(l+1) /Kδl ) → H. It follows that δ δ =K Kδ(l+1) /Kδl δ(l+1) /Kδl . δ =g Therefore, since gδl Kδl by the induction hypothesis, δ gδ(l+1) =g δ K /K δ δ(l+1) δl δ =g ◦ gδl Kδ(l+1) /Kδl ◦ gKδl = gKδ(l+1) . If µk = µδk is a sequence with a weak limit µ ∈ N , which exists because the hulls Ktδ are uniformly bounded for each t < ∞, then Proposition 9 implies that gtk converges to gt , the solution of the Loewner equation driven by µ. Since f k → f locally uniformly in space and time, it follows that Ktk → Kt in the Carath´eodory topology for every t. Hence, gt solves the Loewner equation driven by µ, and gt : H\Kt → H. The last part of Theorem 8 which remains to be proven is the convergence fact. We need to prove that convergence of a family of hulls implies convergence of the corresponding measures. Proposition 11. Suppose that µn ∈ N , and that gtn : H\Ktn → H is the associated solution of the chordal Loewner equation. Then if Ktn → Kt in the Carath´eodory topology for some increasing family of hulls Kt and the family {Ktn }n∈N is uniformly bounded for each t, then there is a measure µ ∈ N so that (Kt ) are the hulls associated with µ and µn → µ in N . Proof. By Proposition 9, we know that there is a measure µ which drives the family of 123 hulls Kt . Let µ ˜ be any subsequential limit of µn , which exists because the hulls Ktn are n bounded for each t < ∞. Also by Proposition 9, µnk → µ ˜ implies that Kt k converges in n the Carath´eodory topology to the hulls driven by µ ˜. Since we know that Kt k converges to Kt , and the hulls are uniquely determined by the measures, it follows that µ ˜ = µ. Since µ ˜ was an arbitrary subsequential limit, it follows that µn → µ in N . 124 BIBLIOGRAPHY 125 BIBLIOGRAPHY [1] L.V. Ahlfors (1973).Conformal invariants: McGraw-Hill Book Co., New York. topics in geometric function theory. [2] Alberts, T., Kozdron, M., Lawler, G. The Green’s function for the radial SchrammLoewner evolution. J. Phys. A: Math. Theor., volume 45, number 49, paper 494015, 2012 [3] Beffara, Vincent. The dimension of the SLE curves. Ann. Probab. Volume 36, Number 4 (2008), 1421-1452. [4] Berestyki, N., Jackson, H. The Rohde-Schramm theorem, via the Gaussian free field. arXiv:1703.00682 [5] Field, Laurence S. Two sided radial SLE and length-biased chordal SLE. arXiv:1601.03374 [6] Jones, Peter and Smirnov, Stanislav. Removability theorems for Sobolov functions and quasiconformal maps. Ark. Mat. Volume 38, Number 2 (2000), 263-279. [7] Lawler, Gregory F. Conformally invariant processes in the plane. Mathematical Surveys and Monographs, 2005. [8] Lawler G.F. (2009) Multifractal Analysis of the Reverse Flow for the Schramm-Loewner Evolution. In: Bandt C., Zhle M., Mrters P. (eds) Fractal Geometry and Stochastics IV. Progress in Probability, vol 61. Birkhuser Basel [9] Lawler, Gregory F.; Rezaei, M (2015). Minkowski content and natural parametrization for the Schramm-Loewner evolution. Annals of Probability 2015, Vol. 43, No. 3, 10821120 [10] Lawler, Gregory F.; Rezaei, M (2015). Up-to-constant bounds on the two-point Green’s function for SLE curves. Electron. Commun. Probab. Volume 20 (2015), paper no. 45, 13 pp. [11] Lawler, Gregory F.; Schramm, Oded; Werner, Wendelin. Conformal restriction: the chordal case.... [12] Lawler, Gregory F.; Sheffield, S. A natural parametrization for the Schramm-Loewner evolution. Annals of Probab. 39, 1896-1937. 126 [13] Lawler, Gregory F.; Schramm, Oded; Werner, Wendelin. Conformal invariance of planar loop-erased random walks and uniform spanning trees. Ann. Probab. 32 (2004), no. 1B, 939–995. [14] Lawler, Gregory F.; Viklund, Fredrik (2016). Convergence of loop-erased random walk in the natural parametrization. arXiv:1603.05203v2 [15] Lawler, Gregory F.; Wener, Wendelin. The Brownian loop soup. [16] Lawler, Gregory F.; Werness, B (2013). Multi-point Green’s function for SLE and an estimate of Beffara. Annals of Probab. Vol. 41, No. 3A (2013), 15131555 [17] Lawler, Gregory F.; Zhou, W (2013). SLE curves and natural parametrization. Ann. Probab. Volume 41, Number 3A (2013), 1556-1584. [18] L¨owner, K.. ”Untersuchungen ber schlichte konforme Abbildungen des Einheitskreises. I..” Mathematische Annalen 89 (1923): 103-121. [19] Miller, Jason and Sheffield, Scott. Quantum Loewner Evolution. arXiv:1312.5745 [20] Miller, Jason and Sheffield, Scott. Imaginary geometry III: reversibility of SLEκ for κ ∈ (4, 8) arXiv:1201.1498 [21] Miller, Jason and Sheffield, Scott. Imaginary Geometry IV: interior rays, whole-plane reversibility, and space-filling trees, arXiv:1302.4738. [22] Rezaei, Mohammad (2012). Hausdorff measure of SLE curves. arXiv:1212.5847 [23] Rezaei, M.; Zhan, D. Higher moments of the natural parametrization for SLE curves. Ann. Inst. H. Poincar Probab. Statist. Volume 53, Number 1 (2017), 182-199. [24] Rezaei, M.; Zhan, D. Green’s function for chordal SLE curves. In preprint (2016), arXiv:1607.03840v2 [25] Rhode, S., Tran, H., and Zinsmeister, M. The Loewner Equation and Lipschitz graphs. Preprint, 2013; [26] Rhode, Steffon and Zhan, Dapeng. Backward SLE and the symmetry of the welding. Probability Theory and Related Fields, 164(3):815-863, 2016. [27] Rhode, Steffon and Schramm, Oded. Basic properties of SLE. Annals Math. 161 (2005) 879-920 [28] Rudin, Walter. Real and Complex Analysis. New York: McGraw-Hill, 1987. 127 [29] O. Schramm (2000), Scaling limits of loop-erased random walks and uniform spanning trees, Israel J. Math. 118, 221288. [30] Schramm, O.; Sheffield, S (2005). Harmonic explorer and its convergence to SLE4 , Annals of Probab. 33, 2127-2148. [31] Schramm, Oded and Wilson, David. SLE coordinate changes. New York J. Math.11:659-669, 2005 [32] Sheffield, Scott. Exploration trees and conformal loop ensembles [33] Sheffield, Scott. Conformal weldings of random surfaces: SLE and the quantum gravity zipper. Ann. Probab. Volume 44, Number 5 (2016), 3474-3545. [34] Smirnov, S (2001). Critical percolation in the plane: Conformal invariance, Cardy’s formula, scaling limits. C.R. Acad. Sci. Paris Ser. I Math. 33 239-244. [35] Smirnov, S (2009). Conformal invariance in random cluster models. I. Holomorphic fermions in the Ising model, Ann of Math. 172 1435-1467 [36] Zhan, Dapeng. The scaling limits of planar LERW in finitely connected domains. Ann. Probab. 36 (2008), no. 2, 467–529. doi:10.1214/07-AOP342. http://projecteuclid.org/euclid.aop/1204306959. [37] Zhan, Dapeng. Decomposition of Schramm-Loewner evolution along its curve. arXiv:1509.05015v2 [38] Zhan, D. Reversibility of chordal SLE. Annals of Probability, 36(4):1472-1494, 2008. [39] Zhan, Dapeng. Ergodicity of http://arxiv.org/pdf/1310.2573.pdf the curve of an SLE curve . arxiv [40] Zhan, Dapeng. Reversibility of Whole-Plane SLE. Probability Theory and Related Fields, 161(3):561-618, 2015. [41] Zhan, Dapeng (2017). SLE Loop Measures. arXiv:1702.08026 128